Sei sulla pagina 1di 100

Contents

1 Oil refinery 1
1.1 Operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Major products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Common process units found in a refinery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.3.1 Flow diagram of typical refinery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.3.2 The crude oil distillation unit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.4 Specialty end products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.5 Siting/locating of petroleum refineries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.6 Safety and environmental concerns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.7 Corrosion problems and prevention . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.8 History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.8.1 Oil refining in the United States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.9 See also . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.10 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.11 External links . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

2 Continuous distillation 14
2.1 Industrial application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.2 Principle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3 Design and operation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.3.1 Column feed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.3.2 Improving separation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.3.3 Overhead system arrangements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.4 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.4.1 Continuous distillation of crude oil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.5 See also . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.6 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.7 External links . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

3 Vacuum distillation 24
3.1 Laboratory-scale applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.1.1 Rotary evaporation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.1.2 Distillation of high-boiling and/or air sensitive materials . . . . . . . . . . . . . . . . . . . 24

i
ii CONTENTS

3.2 Industrial-scale applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26


3.2.1 Vacuum distillation in petroleum refining . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.3 Molecular distillation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.4 Gallery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.5 See also . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.6 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
3.7 External links . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

4 Hydrodesulfurization 31
4.1 History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
4.2 Process chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.3 Process description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.4 Sulfur compounds in refinery HDS feedstocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.4.1 Thiophenes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.5 Catalysts and mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.5.1 Catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.5.2 Supports . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.6 Other uses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.6.1 Hydrodenitrogenation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
4.6.2 Saturation of olefins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.6.3 Hydrogenation in the food industry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.7 See also . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.8 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.9 External links . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

5 Catalytic reforming 37
5.1 History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
5.2 Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
5.2.1 Typical naphtha feedstocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
5.2.2 The reaction chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
5.3 Process description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
5.4 Catalysts and mechanisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
5.5 Economics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
5.6 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
5.7 External links . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

6 Fluid catalytic cracking 44


6.1 Flow diagram and process description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
6.1.1 Reactor and Regenerator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
6.1.2 Distillation column . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
6.1.3 Regenerator flue gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
6.2 Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
CONTENTS iii

6.2.1 Catalysts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
6.3 History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
6.4 See also . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
6.5 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
6.6 External links . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

7 Cracking (chemistry) 54
7.1 History and patents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
7.2 Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
7.2.1 Initiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
7.2.2 Hydrogen abstraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
7.2.3 Radical decomposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
7.2.4 Radical addition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
7.2.5 Termination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
7.2.6 Example: cracking butane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
7.3 Cracking methodologies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
7.3.1 Thermal methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
7.3.2 Catalytic methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
7.4 See also . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
7.5 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
7.6 External links . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

8 Visbreaker 61
8.1 Process objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
8.2 Technology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
8.2.1 Coil visbreaking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
8.2.2 Soaker visbreaking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
8.2.3 Process options . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
8.2.4 Soaker visbreaking versus coil visbreaking . . . . . . . . . . . . . . . . . . . . . . . . . . 62
8.3 Quality and yields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
8.3.1 Feed quality and product quality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
8.3.2 Yields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
8.3.3 Fuel oil stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
8.4 Economics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
8.4.1 Viscosity blending . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
8.4.2 Example economics for a two-component blend . . . . . . . . . . . . . . . . . . . . . . . 64
8.5 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
8.6 External links . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65

9 Merox 66
9.1 Types of Merox process units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
9.2 Conventional Merox for extracting mercaptans from LPG . . . . . . . . . . . . . . . . . . . . . . 67
iv CONTENTS

9.2.1 Flow diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68


9.3 Conventional Merox for sweetening jet fuel or kerosene . . . . . . . . . . . . . . . . . . . . . . . 68
9.3.1 Flow diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
9.4 See also . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
9.5 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
9.6 External links . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

10 Doctor sweetening process 71


10.1 Chemistry of the process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
10.2 Literature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
10.3 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
10.4 External links . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

11 Coker unit 74
11.1 Types of coker units . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
11.2 See also . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
11.3 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
11.4 External links . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75

12 Alkylation 76
12.1 Alkylating agents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
12.1.1 Nucleophilic alkylating agents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
12.1.2 Electrophilic alkylating agents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
12.1.3 Carbene alkylating agents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
12.2 In biology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
12.3 Oil refining . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
12.4 See also . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
12.5 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
12.6 External links . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

13 Isomerization 80
13.1 Instances of isomerization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
13.2 See also . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
13.3 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

14 Amine gas treating 82


14.1 Description of a typical amine treater . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
14.2 Amines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
14.3 Uses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
14.4 See also . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
14.5 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
14.6 External links . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85

15 Claus process 86
CONTENTS v

15.1 History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
15.2 Process description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
15.2.1 Thermal step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
15.2.2 Catalytic step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
15.2.3 Sub dew point Claus process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
15.3 Process performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
15.4 See also . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
15.5 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
15.6 Text and image sources, contributors, and licenses . . . . . . . . . . . . . . . . . . . . . . . . . . 91
15.6.1 Text . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
15.6.2 Images . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
15.6.3 Content license . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
Chapter 1

Oil refinery

Anacortes Refinery (Tesoro), on the north end of March Point southeast of Anacortes, Washington

An oil refinery or petroleum refinery is an industrial process plant where crude oil is processed and refined into
more useful products such as petroleum naphtha, gasoline, diesel fuel, asphalt base, heating oil, kerosene and liquefied
petroleum gas.[1][2] Oil refineries are typically large, sprawling industrial complexes with extensive piping running
throughout, carrying streams of fluids between large chemical processing units. In many ways, oil refineries use much
of the technology of, and can be thought of, as types of chemical plants. The crude oil feedstock has typically been
processed by an oil production plant. There is usually an oil depot (tank farm) at or near an oil refinery for the storage
of incoming crude oil feedstock as well as bulk liquid products.
An oil refinery is considered an essential part of the downstream side of the petroleum industry.

1
2 CHAPTER 1. OIL REFINERY

Crude oil is separated into fractions by fractional distillation. The fractions at the top of the fractionating column have lower boiling
points than the fractions at the bottom. The heavy bottom fractions are often cracked into lighter, more useful products. All of the
fractions are processed further in other refining units.

1.1 Operation

Raw or unprocessed crude oil is not generally useful in industrial applications, although “light, sweet” (low viscosity,
low sulfur) crude oil has been used directly as a burner fuel to produce steam for the propulsion of seagoing vessels.
The lighter elements, however, form explosive vapors in the fuel tanks and are therefore hazardous, especially in
warships. Instead, the hundreds of different hydrocarbon molecules in crude oil are separated in a refinery into
components which can be used as fuels, lubricants, and as feedstocks in petrochemical processes that manufacture
such products as plastics, detergents, solvents, elastomers and fibers such as nylon and polyesters.
Petroleum fossil fuels are burned in internal combustion engines to provide power for ships, automobiles, aircraft
engines, lawn mowers, chainsaws, and other machines. Different boiling points allow the hydrocarbons to be separated
by distillation. Since the lighter liquid products are in great demand for use in internal combustion engines, a modern
refinery will convert heavy hydrocarbons and lighter gaseous elements into these higher value products.
Oil can be used in a variety of ways because it contains hydrocarbons of varying molecular masses, forms and lengths
such as paraffins, aromatics, naphthenes (or cycloalkanes), alkenes, dienes, and alkynes. While the molecules in crude
oil include different atoms such as sulfur and nitrogen, the hydrocarbons are the most common form of molecules,
1.2. MAJOR PRODUCTS 3

The oil refinery in Haifa, Israel is capable of processing about 9 million tons (66 million barrels) of crude oil a year. Its two cooling
towers are landmarks of the city’s skyline.

which are molecules of varying lengths and complexity made of hydrogen and carbon atoms, and a small number of
oxygen atoms. The differences in the structure of these molecules account for their varying physical and chemical
properties, and it is this variety that makes crude oil useful in a broad range of several applications.
Once separated and purified of any contaminants and impurities, the fuel or lubricant can be sold without further
processing. Smaller molecules such as isobutane and propylene or butylenes can be recombined to meet specific
octane requirements by processes such as alkylation, or less commonly, dimerization. Octane grade of gasoline can
also be improved by catalytic reforming, which involves removing hydrogen from hydrocarbons producing compounds
with higher octane ratings such as aromatics. Intermediate products such as gasoils can even be reprocessed to break a
heavy, long-chained oil into a lighter short-chained one, by various forms of cracking such as fluid catalytic cracking,
thermal cracking, and hydrocracking. The final step in gasoline production is the blending of fuels with different
octane ratings, vapor pressures, and other properties to meet product specifications.
Oil refineries are large scale plants, processing about a hundred thousand to several hundred thousand barrels of crude
oil a day. Because of the high capacity, many of the units operate continuously, as opposed to processing in batches,
at steady state or nearly steady state for months to years. The high capacity also makes process optimization and
advanced process control very desirable.

1.2 Major products


Petroleum products are usually grouped into three categories: light distillates (LPG, gasoline, naphtha), middle dis-
tillates (kerosene, diesel), heavy distillates and residuum (heavy fuel oil, lubricating oils, wax, asphalt). This classifi-
cation is based on the way crude oil is distilled and separated into fractions (called distillates and residuum) as in the
above drawing.[2]

• Liquified petroleum gas (LPG)


4 CHAPTER 1. OIL REFINERY

• Gasoline (also known as petrol)


• Naphtha
• Kerosene and related jet aircraft fuels
• Diesel fuel
• Fuel oils
• Lubricating oils
• Paraffin wax
• Asphalt and tar
• Petroleum coke
• Sulfur

Oil refineries also produce various intermediate products such as hydrogen, light hydrocarbons, reformate and pyrolysis
gasoline. These are not usually transported but instead are blended or processed further on-site. Chemical plants are
thus often adjacent to oil refineries. For example, light hydrocarbons are steam-cracked in an ethylene plant, and the
produced ethylene is polymerized to produce polyethene.

1.3 Common process units found in a refinery

Storage tanks and towers at Shell Puget Sound Refinery (Shell Oil Company), Anacortes, Washington

• Desalter unit washes out salt from the crude oil before it enters the atmospheric distillation unit.
• Atmospheric distillation unit distills crude oil into fractions. See continuous distillation.
1.3. COMMON PROCESS UNITS FOUND IN A REFINERY 5

• Vacuum distillation unit further distills residual bottoms after atmospheric distillation.

• Naphtha hydrotreater unit uses hydrogen to desulfurize naphtha from atmospheric distillation. Must hydrotreat
the naphtha before sending to a catalytic reformer unit.

• Catalytic reformer unit is used to convert the naphtha-boiling range molecules into higher octane reformate
(reformer product). The reformate has higher content of aromatics and cyclic hydrocarbons). An important
byproduct of a reformer is hydrogen released during the catalyst reaction. The hydrogen is used either in the
hydrotreaters or the hydrocracker.

• Distillate hydrotreater desulfurizes distillates (such as diesel) after atmospheric distillation.

• Fluid Catalytic Cracker (FCC) unit upgrades heavier fractions into lighter, more valuable products.

• Hydrocracker unit uses hydrogen to upgrade heavier fractions into lighter, more valuable products.

• Visbreaking unit upgrades heavy residual oils by thermally cracking them into lighter, more valuable reduced
viscosity products.

• Merox unit treats LPG, kerosene or jet fuel by oxidizing mercaptans to organic disulfides.

• Alternative processes for removing mercaptans are known, e.g. doctor sweetening process and caustic washing.

• Coking units (delayed coking, fluid coker, and flexicoker) process very heavy residual oils into gasoline and
diesel fuel, leaving petroleum coke as a residual product.

• Alkylation unit uses sulfuric acid or hydrofluoric acid to produce high-octane components for gasoline blending.

• Dimerization unit converts olefins into higher-octane gasoline blending components. For example, butenes can
be dimerized into isooctene which may subsequently be hydrogenated to form isooctane. There are also other
uses for dimerization. Gasoline produced through dimerization is highly unsaturated and very reactive. It tends
spontaneously to form gums. For this reason the effluent from the dimerization need to be blended into the
finished gasoline pool immediately or hydrogenated.

• Isomerization unit converts linear molecules to higher-octane branched molecules for blending into gasoline or
feed to alkylation units.

• Steam reforming unit produces hydrogen for the hydrotreaters or hydrocracker.

• Liquified gas storage vessels store propane and similar gaseous fuels at pressure sufficient to maintain them in
liquid form. These are usually spherical vessels or “bullets” (i.e., horizontal vessels with rounded ends).

• Storage tanks store crude oil and finished products, usually cylindrical, with some sort of vapor emission control
and surrounded by an earthen berm to contain spills.

• Amine gas treater, Claus unit, and tail gas treatment convert hydrogen sulfide from hydrodesulfurization into
elemental sulfur.

• Utility units such as cooling towers circulate cooling water, boiler plants generates steam, and instrument air
systems include pneumatically operated control valves and an electrical substation.

• Wastewater collection and treating systems consist of API separators, dissolved air flotation (DAF) units and
further treatment units such as an activated sludge biotreater to make water suitable for reuse or for disposal.[3]

• Solvent refining units use solvent such as cresol or furfural to remove unwanted, mainly aromatics from lubri-
cating oil stock or diesel stock.

• Solvent dewaxing units remove the heavy waxy constituents petrolatum from vacuum distillation products.
6 CHAPTER 1. OIL REFINERY

Schematic flow diagram of a typical oil refinery

1.3.1 Flow diagram of typical refinery

The image below is a schematic flow diagram of a typical oil refinery[4] that depicts the various unit processes and
the flow of intermediate product streams that occurs between the inlet crude oil feedstock and the final end products.
The diagram depicts only one of the literally hundreds of different oil refinery configurations. The diagram also does
not include any of the usual refinery facilities providing utilities such as steam, cooling water, and electric power as
well as storage tanks for crude oil feedstock and for intermediate products and end products.[1][5][6][7]
There are many process configurations other than that depicted above. For example, the vacuum distillation unit
may also produce fractions that can be refined into endproducts such as: spindle oil used in the textile industry, light
machinery oil, motor oil, and various waxes.
1.3. COMMON PROCESS UNITS FOUND IN A REFINERY 7

1.3.2 The crude oil distillation unit

The crude oil distillation unit (CDU) is the first processing unit in virtually all petroleum refineries. The CDU distills
the incoming crude oil into various fractions of different boiling ranges, each of which are then processed further
in the other refinery processing units. The CDU is often referred to as the atmospheric distillation unit because it
operates at slightly above atmospheric pressure.[1][2][8]
Below is a schematic flow diagram of a typical crude oil distillation unit. The incoming crude oil is preheated by
exchanging heat with some of the hot, distilled fractions and other streams. It is then desalted to remove inorganic
salts (primarily sodium chloride).
Following the desalter, the crude oil is further heated by exchanging heat with some of the hot, distilled fractions and
other streams. It is then heated in a fuel-fired furnace (fired heater) to a temperature of about 398 °C and routed into
the bottom of the distillation unit.
The cooling and condensing of the distillation tower overhead is provided partially by exchanging heat with the
incoming crude oil and partially by either an air-cooled or water-cooled condenser. Additional heat is removed from
the distillation column by a pumparound system as shown in the diagram below.
As shown in the flow diagram, the overhead distillate fraction from the distillation column is naphtha. The fractions
removed from the side of the distillation column at various points between the column top and bottom are called
sidecuts. Each of the sidecuts (i.e., the kerosene, light gas oil and heavy gas oil) is cooled by exchanging heat with
the incoming crude oil. All of the fractions (i.e., the overhead naphtha, the sidecuts and the bottom residue) are sent
to intermediate storage tanks before being processed further.

Schematic flow diagram of a typical crude oil distillation unit as used in petroleum crude oil refineries.
8 CHAPTER 1. OIL REFINERY

1.4 Specialty end products


These require blending various feedstocks, mixing appropriate additives, providing short term storage, and preparation
for bulk loading to trucks, barges, product ships, and railcars:

• Gaseous fuels such as propane, stored and shipped in liquid form under pressure in specialized railcars to
distributors.

• Lubricants (produces light machine oils, motor oils, and greases, adding viscosity stabilizers as required), usu-
ally shipped in bulk to an offsite packaging plant.

• Wax (paraffin), used in the packaging of frozen foods, among others. May be shipped in bulk to a site to
prepare as packaged blocks.

• Sulfur (or sulfuric acid), byproducts of sulfur removal from petroleum which may have up to a couple percent
sulfur as organic sulfur-containing compounds. Sulfur and sulfuric acid are useful industrial materials. Sulfuric
acid is usually prepared and shipped as the acid precursor oleum.

• Bulk tar shipping for offsite unit packaging for use in tar-and-gravel roofing.

• Asphalt unit. Prepares bulk asphalt for shipment.

• Petroleum coke, used in specialty carbon products or as solid fuel.

• Petrochemicals or petrochemical feedstocks, which are often sent to petrochemical plants for further processing
in a variety of ways. The petrochemicals may be olefins or their precursors, or various types of aromatic
petrochemicals.

1.5 Siting/locating of petroleum refineries


A party searching for a site to construct a refinery or a chemical plant needs to consider the following issues:

• The site has to be reasonably far from residential areas.

• Infrastructure should be available for supply of raw materials and shipment of products to markets.

• Energy to operate the plant should be available.

• Facilities should be available for waste disposal.

Refineries which use a large amount of steam and cooling water need to have an abundant source of water. Oil
refineries therefore are often located nearby navigable rivers or on a sea shore, nearby a port. Such location also gives
access to transportation by river or by sea. The advantages of transporting crude oil by pipeline are evident, and oil
companies often transport a large volume of fuel to distribution terminals by pipeline. Pipeline may not be practical
for products with small output, and rail cars, road tankers, and barges are used.
Petrochemical plants and solvent manufacturing (fine fractionating) plants need spaces for further processing of a
large volume of refinery products for further processing, or to mix chemical additives with a product at source rather
than at blending terminals.

1.6 Safety and environmental concerns


The refining process releases a number of different chemicals into the atmosphere (see AP 42 Compilation of Air
Pollutant Emission Factors) and a notable odor normally accompanies the presence of a refinery. Aside from air
pollution impacts there are also wastewater concerns,[3] risks of industrial accidents such as fire and explosion, and
noise health effects due to industrial noise.
1.7. CORROSION PROBLEMS AND PREVENTION 9

Fire-extinguishing operations after the Texas City refinery explosion.

Many governments worldwide have mandated restrictions on contaminants that refineries release, and most refineries
have installed the equipment needed to comply with the requirements of the pertinent environmental protection
regulatory agencies. In the United States, there is strong pressure to prevent the development of new refineries, and
no major refinery has been built in the country since Marathon’s Garyville, Louisiana facility in 1976. However,
many existing refineries have been expanded during that time. Environmental restrictions and pressure to prevent
construction of new refineries may have also contributed to rising fuel prices in the United States.[9] Additionally,
many refineries (more than 100 since the 1980s) have closed due to obsolescence and/or merger activity within the
industry itself.
Environmental and safety concerns mean that oil refineries are sometimes located some distance away from major
urban areas. Nevertheless, there are many instances where refinery operations are close to populated areas and pose
health risks In California's Contra Costa County and Solano County, a shoreline necklace of refineries, built in the
early 20th century before this area was populated, and associated chemical plants are adjacent to urban areas in
Richmond, Martinez, Pacheco, Concord, Pittsburg, Vallejo and Benicia, with occasional accidental events that require
"shelter in place" orders to the adjacent populations.

1.7 Corrosion problems and prevention


Petroleum refineries run as efficiently as possible to reduce costs. One major factor that decreases efficiency is corro-
sion of the metallic components found throughout refining process. Corrosion causes the failure of equipment items
as well as dictating the maintenance schedule of the refinery, during which part or all of the refinery must be shut
down. The corrosion-related direct costs in the U.S. petroleum industry as of 1996 was estimated as US$3.7 billion
per year.[10][11]
Corrosion occurs in various forms in the refining process, such as pitting corrosion from water droplets, embrittlement
from hydrogen, and stress corrosion cracking from sulfide attack.[12] From a materials standpoint, carbon steel is used
for upwards of 80 per cent of refinery components, which is beneficial due to its low cost. Carbon steel is resistant
to the most common forms of corrosion, particularly from hydrocarbon impurities at temperatures below 205 °C,
but other corrosive chemicals and environments prevent its use everywhere. Common replacement materials are low
alloy steels containing chromium and molybdenum, with stainless steels containing more chromium dealing with more
10 CHAPTER 1. OIL REFINERY

Refinery of Slovnaft in Bratislava.

corrosive environments. More expensive materials commonly used are nickel, titanium, and copper alloys. These are
primarily saved for the most problematic areas where extremely high temperatures and/or very corrosive chemicals
are present.[13]
Corrosion is fought by a complex system of monitoring, preventative repairs and careful use of materials. Monitoring
methods include both off-line checks taken during maintenance and on-line monitoring. Off-line checks measure cor-
rosion after it has occurred, telling the engineer when equipment must be replaced based on the historical information
he has collected. This is referred to as preventative management.
On-line systems are a more modern development, and are revolutionizing the way corrosion is approached. There are
several types of on-line corrosion monitoring technologies such as linear polarization resistance, electrochemical noise
and electrical resistance. On-Line monitoring has generally had slow reporting rates in the past (minutes or hours)
and been limited by process conditions and sources of error but newer technologies can report rates up to twice
per minute with much higher accuracy (referred to as real-time monitoring). This allows process engineers to treat
corrosion as another process variable that can be optimized in the system. Immediate responses to process changes
allow the control of corrosion mechanisms, so they can be minimized while also maximizing production output.[14] In
an ideal situation having on-line corrosion information that is accurate and real-time will allow conditions that cause
high corrosion rates to be identified and reduced. This is known as predictive management.
Materials methods include selecting the proper material for the application. In areas of minimal corrosion, cheap
materials are preferable, but when bad corrosion can occur, more expensive but longer lasting materials should be used.
Other materials methods come in the form of protective barriers between corrosive substances and the equipment
metals. These can be either a lining of refractory material such as standard Portland cement or other special acid-
resistant cements that are shot onto the inner surface of the vessel. Also available are thin overlays of more expensive
metals that protect cheaper metal against corrosion without requiring lots of material.[15]

1.8 History
The world’s first refinery opened at Ploiești, Romania, in 1856-1857,[16] with United States investment. After being
taken over by Nazi Germany, the Ploiești refineries were bombed in Operation Tidal Wave by the Allies during the
Oil Campaign of World War II.
1.9. SEE ALSO 11

At one point, the refinery in Ras Tanura, Saudi Arabia owned by Saudi Aramco was claimed to be the largest oil
refinery in the world. For most of the 20th century, the largest refinery was the Abadan Refinery in Iran. This refinery
suffered extensive damage during the Iran-Iraq war. The world’s largest refinery complex is the Jamnagar Refinery
Complex, consisting of two refineries side by side operated by Reliance Industries Limited in Jamnagar, India with a
combined production capacity of 1,240,000 barrels per day (197,000 m3 /d). PDVSA's Paraguaná Refinery Complex
in Paraguaná Peninsula, Venezuela with a capacity of 956,000 bbl/d (152,000 m3 /d) and SK Energy's Ulsan in South
Korea with 840,000 bbl/d (134,000 m3 /d) are the second and third largest, respectively.

1.8.1 Oil refining in the United States


In the 19th century, refineries in the U.S. processed crude oil primarily to recover the kerosene. There was no
market for the more volatile fraction, including gasoline, which was considered waste and was often dumped directly
into the nearest river. The invention of the automobile shifted the demand to gasoline and diesel, which remain the
primary refined products today. Today, national and state legislation requires refineries to meet stringent air and water
cleanliness standards. In fact, oil companies in the U.S. perceive obtaining a permit to build a modern refinery to
be so difficult and costly that no new refineries were built (though many have been expanded) in the U.S. from 1976
until 2014, when the small Dakota Prairie Refinery in North Dakota is set to begin operation.[17] More than half the
refineries that existed in 1981 are now closed due to low utilization rates and accelerating mergers.[18] As a result
of these closures total US refinery capacity fell between 1981 to 1995, though the operating capacity stayed fairly
constant in that time period at around 15,000,000 barrels per day (2,400,000 m3 /d).[19] Increases in facility size and
improvements in efficiencies have offset much of the lost physical capacity of the industry. In 1982 (the earliest data
provided), the United States operated 301 refineries with a combined capacity of 17.9 million barrels (2,850,000 m3 )
of crude oil each calendar day. In 2010, there were 149 operable U.S. refineries with a combined capacity of 17.6
million barrels (2,800,000 m3 ) per calendar day.[20]
In 2009 through 2010, as revenue streams in the oil business dried up and profitability of oil refineries fell due to lower
demand for product and high reserves of supply preceding the economic recession, oil companies began to close or
sell refineries.

ExxonMobil oil refinery in Baton Rouge (the second-largest in the United States)

1.9 See also


• Acid gas

• H-Bio

• AP 42 Compilation of Air Pollutant Emission Factors

• API oil-water separator

• Ethanol fuel

• Butanol fuel

• Gas flare

• Industrial wastewater treatment

• K factor crude oil refining


12 CHAPTER 1. OIL REFINERY

• List of oil refineries

• Natural-gas processing

• Nelson complexity index

• Sour gas

1.10 References
[1] Gary, J.H. and Handwerk, G.E. (1984). Petroleum Refining Technology and Economics (2nd ed.). Marcel Dekker, Inc.
ISBN 0-8247-7150-8.

[2] Leffler, W.L. (1985). Petroleum refining for the nontechnical person (2nd ed.). PennWell Books. ISBN 0-87814-280-0.

[3] Beychok, Milton R. (1967). Aqueous Wastes from Petroleum and Petrochemical Plants (1st ed.). John Wiley & Sons.
LCCN 67019834.

[4] Crude Oil Solids Removal

[5] Guide to Refining from Chevron Oil's website

[6] Refinery flowchart from Universal Oil Products' website

[7] An example flowchart of fractions from crude oil at a refinery

[8] Kister, Henry Z. (1992). Distillation Design (1st ed.). McGraw-Hill. ISBN 978-0-07-034909-4.

[9] By Steve Hargreaves, CNNMoney.com staff writer (2007-04-17). “Behind high gas prices: The refinery crunch”. Money.cnn.com.
Retrieved 2011-11-05.

[10] Corrosion Costs and Preventive Strategies in the United States, a publication of NACE International.

[11] R.D. Kane, Corrosion in Petroleum Refining and Petrochemical Operations, Corrosion: Environments and Industries, Vol
13C, ASM Handbook, ASM International, 2006, p 967–1014.

[12] E.N. Skinner, J.F. Mason, and J.J. Moran, High Temperature Corrosion in Refinery and Petrochemical Service, Corrosion,
Vol 16 (No. 12), 1960, p 593t–600t.

[13] E.L. Hildebrand, Materials Selection for Petroleum Refineries and Petrochemical Plants, Mater. Prot. Perform., Vol 11
(No. 7), 1972, p19–22.

[14] R.D. Kane, D.C. Eden, and D.A. Eden, Innovative Solutions Integrate Corrosion Monitoring with Process Control, Mater.
Perform., Feb 2005, p 36–41.

[15] W.A. McGill and M.J. Weinbaum, Aluminum-Diffused Steel Lasts Longer, Oil Gas J., Vol 70, Oct 9, 1972, p 66–69.

[16] “WORLD EVENTS: 1844-1856”. PBS.org. Retrieved 2009-04-22. world’s first oil refinery

[17] “North Dakota Builds A Refinery, First In The U.S. Since '76”. Investors Business Daily. April 11, 2013. Retrieved August
24, 2014.

[18] White Paper on Refining Capacity, Federal Trade Commission, April, 2007.

[19] “U. S. Operating Crude Oil Distillation Capacity (Thousand Barrels per Day)". Eia.doe.gov. 2011-07-28. Retrieved
2011-11-05.

[20] “2011 The U.S. Petroleum Industry: Statistics & Definitions” (PDF). Retrieved 2011-11-05.
1.11. EXTERNAL LINKS 13

1.11 External links


• Interactive map of UK refineries

• Searchable United States Refinery Map


• Complete, detailed refinery description

• Ecomuseum Bergslagen - history of Oljeön, Sweden

• Fueling Profits: Report on Industry Consolidation (publication of the Consumer Federation of America)
• Price Spikes, Excess Profits and Excuses (publication of the Consumer Federation of America)

• Basics of Oil Refining Overview of crude oil refining process


• Refining NZ Learning Centre Oil Refinery Process Animations,Videos & 360 Degree Views
Chapter 2

Continuous distillation

Continuous distillation, a form of distillation, is an ongoing separation in which a mixture is continuously (without
interruption) fed into the process and separated fractions are removed continuously as output streams. A distillation
is the separation or partial separation of a liquid feed mixture into components or fractions by selective boiling (or
evaporation) and condensation. A distillation produces at least two output fractions. These fractions include at least
one volatile distillate fraction, which has boiled and been separately captured as a vapor condensed to a liquid, and
practically always a bottoms (or residuum) fraction, which is the least volatile residue that has not been separately
captured as a condensed vapor.
An alternative to continuous distillation is batch distillation, where the mixture is added to the unit at the start of the
distillation, distillate fractions are taken out sequentially in time (one after another) during the distillation, and the
remaining bottoms fraction is removed at the end. Because each of the distillate fractions are taken out at different
times, only one distillate exit point (location) is needed for a batch distillation and the distillate can just be switched
to a different receiver, a fraction-collecting container. Batch distillation is often used when smaller quantities are
distilled. In a continuous distillation, each of the fraction streams is taken simultaneously throughout operation;
therefore, a separate exit point is needed for each fraction. In practice when there are multiple distillate fractions,
each of the distillate exit points are located at different heights on a fractionating column. The bottoms fraction can
be taken from the bottom of the distillation column or unit, but is often taken from a reboiler connected to the bottom
of the column.
Each fraction may contain one or more components (types of chemical compounds). When distilling crude oil or
a similar feedstock, each fraction contains many components of similar volatility and other properties. Although it
is possible to run a small-scale or laboratory continuous distillation, most often continuous distillation is used in a
large-scale industrial process.

2.1 Industrial application

Distillation is one of the unit operations of chemical engineering.[1][2] Continuous distillation is used widely in the
chemical process industries where large quantities of liquids have to be distilled.[3][4][5] Such industries are the natural
gas processing, petrochemical production, coal tar processing, liquor production, liquified air separation, hydrocarbon
solvents production and similar industries, but it finds its widest application in petroleum refineries. In such refineries,
the crude oil feedstock is a very complex multicomponent mixture that must be separated and yields of pure chemical
compounds are not expected, only groups of compounds within a relatively small range of boiling points, which are
called fractions. These fractions are the origin of the term fractional distillation or fractionation. It is often not
worthwhile separating the components in these fractions any further based on product requirements and economics.
Industrial distillation is typically performed in large, vertical cylindrical columns (as shown in images 1 and 2) known
as “distillation towers” or “distillation columns” with diameters ranging from about 65 centimeters to 11 meters and
heights ranging from about 6 meters to 60 meters or more.

14
2.2. PRINCIPLE 15

Image 1: Typical industrial distillation towers

2.2 Principle

Main article: Distillation


The principle for continuous distillation is the same as for normal distillation: when a liquid mixture is heated so
that it boils, the composition of the vapor above the liquid differs from the liquid composition. If this vapor is then
separated and condensed into a liquid, it becomes richer in the lower boiling component(s) of the original mixture.
16 CHAPTER 2. CONTINUOUS DISTILLATION

This is what happens in a continuous distillation column. A mixture is heated up, and routed into the distillation
column. On entering the column, the feed starts flowing down but part of it, the component(s) with lower boiling
point(s), vaporizes and rises. However, as it rises, it cools and while part of it continues up as vapor, some of it
(enriched in the less volatile component) begins to descend again.
Image 3 depicts a simple continuous fractional distillation tower for separating a feed stream into two fractions, an
overhead distillate product and a bottoms product. The “lightest” products (those with the lowest boiling point or
highest volatility) exit from the top of the columns and the “heaviest” products (the bottoms, those with the highest
boiling point) exit from the bottom of the column. The overhead stream may be cooled and condensed using a water-
cooled or air-cooled condenser. The bottoms reboiler may be a steam-heated or hot oil-heated heat exchanger, or
even a gas or oil-fired furnace.
In a continuous distillation, the system is kept in a steady state or approximate steady state. Steady state means that
quantities related to the process do not change as time passes during operation. Such constant quantities include feed
input rate, output stream rates, heating and cooling rates, reflux ratio, and temperatures, pressures, and compositions
at every point (location). Unless the process is disturbed due to changes in feed, heating, ambient temperature, or
condensing, steady state is normally maintained. This is also the main attraction of continuous distillation, apart
from the minimum amount of (easily instrumentable) surveillance; if the feed rate and feed composition are kept
constant, product rate and quality are also constant. Even when a variation in conditions occurs, modern process
control methods are commonly able to gradually return the continuous process to another steady state again.
Since a continuous distillation unit is fed constantly with a feed mixture and not filled all at once like a batch distillation,
a continuous distillation unit does not need a sizable distillation pot, vessel, or reservoir for a batch fill. Instead, the
mixture can be fed directly into the column, where the actual separation occurs. The height of the feed point along
the column can vary on the situation and is designed so as to provide optimal results. See McCabe–Thiele method.
A continuous distillation is often a fractional distillation and can be a vacuum distillation or a steam distillation.

2.3 Design and operation


Design and operation of a distillation column depends on the feed and desired products. Given a simple, binary
component feed, analytical methods such as the McCabe–Thiele method[5][6][7] or the Fenske equation[5] can be used
to assist in the design. For a multi-component feed, computerized simulation models are used both for design and
subsequently in operation of the column as well. Modeling is also used to optimize already erected columns for the
distillation of mixtures other than those the distillation equipment was originally designed for.
When a continuous distillation column is in operation, it has to be closely monitored for changes in feed composition,
operating temperature and product composition. Many of these tasks are performed using advanced computer control
equipment.

2.3.1 Column feed

The column can be fed in different ways. If the feed is from a source at a pressure higher than the distillation column
pressure, it is simply piped into the column. Otherwise, the feed is pumped or compressed into the column. The
feed may be a superheated vapor, a saturated vapor, a partially vaporized liquid-vapor mixture, a saturated liquid
(i.e., liquid at its boiling point at the column’s pressure), or a sub-cooled liquid. If the feed is a liquid at a much
higher pressure than the column pressure and flows through a pressure let-down valve just ahead of the column, it
will immediately expand and undergo a partial flash vaporization resulting in a liquid-vapor mixture as it enters the
distillation column.

2.3.2 Improving separation

Although small size units, mostly made of glass, can be used in laboratories, industrial units are large, vertical, steel
vessels (see images 1 and 2) known as “distillation towers” or “distillation columns”. To improve the separation, the
tower is normally provided inside with horizontal plates or trays as shown in image 5, or the column is packed with
a packing material. To provide the heat required for the vaporization involved in distillation and also to compensate
for heat loss, heat is most often added to the bottom of the column by a reboiler, and the purity of the top product can
2.3. DESIGN AND OPERATION 17

be improved by recycling some of the externally condensed top product liquid as reflux. Depending on their purpose,
distillation columns may have liquid outlets at intervals up the length of the column as shown in image 4.

Reflux

Large-scale industrial fractionation towers use reflux to achieve more efficient separation of products.[3][5] Reflux
refers to the portion of the condensed overhead liquid product from a distillation tower that is returned to the upper
part of the tower as shown in images 3 and 4. Inside the tower, the downflowing reflux liquid provides cooling and
partial condensation of the upflowing vapors, thereby increasing the efficacy of the distillation tower. The more reflux
that is provided, the better is the tower’s separation of the lower boiling from the higher boiling components of the
feed. A balance of heating with a reboiler at the bottom of a column and cooling by condensed reflux at the top
of the column maintains a temperature gradient (or gradual temperature difference) along the height of the column
to provide good conditions for fractionating the feed mixture. Reflux flows at the middle of the tower are called
pumparounds.
Changing the reflux (in combination with changes in feed and product withdrawal) can also be used to improve the
separation properties of a continuous distillation column while in operation (in contrast to adding plates or trays, or
changing the packing, which would, at a minimum, require quite significant downtime).

Plates or trays

Distillation towers (such as in images 3 and 4) use various vapor and liquid contacting methods to provide the required
number of equilibrium stages. Such devices are commonly known as “plates” or “trays”.[8] Each of these plates or trays
is at a different temperature and pressure. The stage at the tower bottom has the highest pressure and temperature.
Progressing upwards in the tower, the pressure and temperature decreases for each succeeding stage. The vapor–liquid
equilibrium for each feed component in the tower reacts in its unique way to the different pressure and temperature
conditions at each of the stages. That means that each component establishes a different concentration in the vapor
and liquid phases at each of the stages, and this results in the separation of the components. Some example trays are
depicted in image 5. A more detailed, expanded image of two trays can be seen in the theoretical plate article. The
reboiler often acts as an additional equilibrium stage.
If each physical tray or plate were 100% efficient, then the number of physical trays needed for a given separation
would equal the number of equilibrium stages or theoretical plates. However, that is very seldom the case. Hence, a
distillation column needs more plates than the required number of theoretical vapor–liquid equilibrium stages.

Packing

Another way of improving the separation in a distillation column is to use a packing material instead of trays. These
offer the advantage of a lower pressure drop across the column (when compared to plates or trays), beneficial when
operating under vacuum. If a distillation tower uses packing instead of trays, the number of necessary theoretical
equilibrium stages is first determined and then the packing height equivalent to a theoretical equilibrium stage, known
as the height equivalent to a theoretical plate (HETP), is also determined. The total packing height required is the
number theoretical stages multiplied by the HETP.
This packing material can either be random dumped packing such as Raschig rings or structured sheet metal. Liquids
tend to wet the surface of the packing and the vapors pass across this wetted surface, where mass transfer takes place.
Unlike conventional tray distillation in which every tray represents a separate point of vapor–liquid equilibrium,
the vapor–liquid equilibrium curve in a packed column is continuous. However, when modeling packed columns it
is useful to compute a number of theoretical plates to denote the separation efficiency of the packed column with
respect to more traditional trays. Differently shaped packings have different surface areas and void space between
packings. Both of these factors affect packing performance.
Another factor in addition to the packing shape and surface area that affects the performance of random or structured
packing is liquid and vapor distribution entering the packed bed. The number of theoretical stages required to make
a given separation is calculated using a specific vapor to liquid ratio. If the liquid and vapor are not evenly distributed
across the superficial tower area as it enters the packed bed, the liquid to vapor ratio will not be correct in the packed
bed and the required separation will not be achieved. The packing will appear to not be working properly. The height
equivalent to a theoretical plate (HETP) will be greater than expected. The problem is not the packing itself but the
mal-distribution of the fluids entering the packed bed. Liquid mal-distribution is more frequently the problem than
18 CHAPTER 2. CONTINUOUS DISTILLATION

vapor. The design of the liquid distributors used to introduce the feed and reflux to a packed bed is critical to making
the packing perform at maximum efficiency. Methods of evaluating the effectiveness of a liquid distributor can be
found in references.[9][10]

2.3.3 Overhead system arrangements


Images 4 and 5 assume an overhead stream that is totally condensed into a liquid product using water or air-cooling.
However, in many cases, the tower overhead is not easily condensed totally and the reflux drum must include a vent
gas outlet stream. In yet other cases, the overhead stream may also contain water vapor because either the feed stream
contains some water or some steam is injected into the distillation tower (which is the case in the crude oil distillation
towers in oil refineries). In those cases, if the distillate product is insoluble in water, the reflux drum may contain a
condensed liquid distillate phase, a condensed water phase and a non-condensible gas phase, which makes it necessary
that the reflux drum also have a water outlet stream.

2.4 Examples

2.4.1 Continuous distillation of crude oil


Petroleum crude oils contain hundreds of different hydrocarbon compounds: paraffins, naphthenes and aromatics as
well as organic sulfur compounds, organic nitrogen compounds and some oxygen containing hydrocarbons such as
phenols. Although crude oils generally do not contain olefins, they are formed in many of the processes used in a
petroleum refinery.[11]
The crude oil fractionator does not produce products having a single boiling point; rather, it produces fractions having
boiling ranges.[11][12] For example, the crude oil fractionator produces an overhead fraction called "naphtha" which
becomes a gasoline component after it is further processed through a catalytic hydrodesulfurizer to remove sulfur and
a catalytic reformer to reform its hydrocarbon molecules into more complex molecules with a higher octane rating
value.
The naphtha cut, as that fraction is called, contains many different hydrocarbon compounds. Therefore it has an
initial boiling point of about 35 °C and a final boiling point of about 200 °C. Each cut produced in the fractionating
columns has a different boiling range. At some distance below the overhead, the next cut is withdrawn from the side
of the column and it is usually the jet fuel cut, also known as a kerosene cut. The boiling range of that cut is from
an initial boiling point of about 150 °C to a final boiling point of about 270 °C, and it also contains many different
hydrocarbons. The next cut further down the tower is the diesel oil cut with a boiling range from about 180 °C to
about 315 °C. The boiling ranges between any cut and the next cut overlap because the distillation separations are
not perfectly sharp. After these come the heavy fuel oil cuts and finally the bottoms product, with very wide boiling
ranges. All these cuts are processed further in subsequent refining processes.

2.5 See also


• Azeotropic distillation

• Extractive distillation

• Fractional distillation

• Fractionating column

• Steam distillation

2.6 References
[1] Editors: Jacqueline I. Kroschwitz and Arza Seidel (2004). Kirk-Othmer Encyclopedia of Chemical Technology (5th ed.).
Hoboken, New Jersey: Wiley-Interscience. ISBN 0-471-48810-0.
2.7. EXTERNAL LINKS 19

[2] McCabe, W., Smith, J. and Harriott, P. (2004). Unit Operations of Chemical Engineering (7th ed.). McGraw Hill. ISBN
0-07-284823-5.

[3] Kister, Henry Z. (1992). Distillation Design (1st ed.). McGraw-Hill. ISBN 0-07-034909-6.

[4] King, C.J. (1980). Separation Processes (2nd ed.). McGraw Hill. ISBN 0-07-034612-7.

[5] Perry, Robert H. and Green, Don W. (1984). Perry’s Chemical Engineers’ Handbook (6th ed.). McGraw-Hill. ISBN
0-07-049479-7.

[6] Beychok, Milton (May 1951). “Algebraic Solution of McCabe-Thiele Diagram”. Chemical Engineering Progress.

[7] Seader, J. D., and Henley, Ernest J. (1998). Separation Process Principles. New York: Wiley. ISBN 0-471-58626-9.

[8] Photographs of bubble cap and other tray types (Website of Raschig Gmbh)

[9] Random Packing, Vapor and Liquid Distribution: Liquid and gas distribution in commercial packed towers, Moore, F.,
Rukovena, F., Chemical Plants & Processing, Edition Europe, August 1987, p. 11-15

[10] Structured Packing, Liquid Distribution: A new method to assess liquid distributor quality, Spiegel, L., Chemical Engineering
and Processing 45 (2006), p. 1011-1017

[11] Gary, J.H. and Handwerk, G.E. (1984). Petroleum Refining Technology and Economics (2nd ed.). Marcel Dekker, Inc.
ISBN 0-8247-7150-8.

[12] Nelson, W.L. (1958). Petroleum Refinery Engineering (4th ed.). McGraw Hill. LCCN 57010913.

2.7 External links


• Distillation Theory by Ivar J. Halvorsen and Sigurd Skogestad, Norwegian University of Science and Technol-
ogy, Norway
• Distillation, An Introduction by Ming Tham, Newcastle University, UK

• Distillation by the Distillation Group, USA

• Distillation Lecture Notes by Prof. Randall M. Price at Christian Brothers University


• Petroleum Distillation by Wayne Pafco

• Distillation simulation software


20 CHAPTER 2. CONTINUOUS DISTILLATION
2.7. EXTERNAL LINKS 21

Image 3: Chemical engineering schematic of Continuous Binary Fractional Distillation tower. A binary distillation separates a feed
mixture stream into two fractions: one distillate and one bottoms fractions.
22 CHAPTER 2. CONTINUOUS DISTILLATION

Image 4: Simplified chemical engineering schematic of Continuous Fractional Distillation tower separating one feed mixture stream
into four distillate and one bottoms fractions
2.7. EXTERNAL LINKS 23
Chapter 3

Vacuum distillation

Vacuum distillation is a method of distillation whereby the pressure above the liquid mixture to be distilled is
reduced to less than its vapor pressure (usually less than atmospheric pressure) causing evaporation of the most
volatile liquid(s) (those with the lowest boiling points).[1] This distillation method works on the principle that boiling
occurs when the vapor pressure of a liquid exceeds the ambient pressure. Vacuum distillation is used with or without
heating the mixture.

3.1 Laboratory-scale applications


Laboratory-scale vacuum distillation is used when liquids to be distilled have high atmospheric boiling points or
chemically change at temperatures near their atmospheric boiling points.[2][3][4] Temperature sensitive materials (such
as beta carotene) also require vacuum distillation to remove solvents from the mixture without damaging the product.
Another reason vacuum distillation is used is that compared to steam distillation there is a lower level of residue build
up. This is important in commercial applications where heat transfer is produced using heat exchangers.
Vacuum distillation is sometimes referred to as low temperature distillation.
There are many laboratory applications for vacuum distillation as well as many types of distillation set-ups and ap-
paratuses.
Safety is an important consideration when using glassware as part of the set-up. All of the glass components should be
carefully examined for scratches and cracks which could result in implosions when the vacuum is applied. Wrapping
as much of the glassware with tape as is practical helps to prevent dangerous scattering of glass shards in the event of
an implosion.

3.1.1 Rotary evaporation

Rotary evaporation[5] is a type of vacuum distillation apparatus used to remove bulk solvents from the liquid being
distilled. It is also used by environmental regulatory agencies for determining the amount of solvents in paint, coatings
and inks.[6]
Rotary evaporation set-ups include an apparatus referred to as a Rotovap which rotates the distillation flask (sometimes
called the still pot) to enhance the distillation. Rotating the flask throws up liquid on the walls of the flask and thus
increases the surface area for evaporation.
Heat is often applied to the rotating distillation flask by partially immersing it in a heated bath of water or oil. Typi-
cally, the vacuum in such systems is generated by a water aspirator or a vacuum pump of some type.

3.1.2 Distillation of high-boiling and/or air sensitive materials

Some compounds have high boiling point temperatures as well as being air sensitive. A simple laboratory vacuum
distillation glassware set-up can be used, in which the vacuum can be replaced with an inert gas after the distillation
is complete.

24
3.1. LABORATORY-SCALE APPLICATIONS 25

Figure 1: At atmospheric pressure, dimethyl sulfoxide boils at 189°C. Under a vacuum, it distills off into the connected receiver at
only 70°C.

However, this is not a completely satisfactory system if it is desired to collect fractions under a reduced pressure.
For better results or for very air sensitive compounds, either a Perkin triangle distillation set-up or a short-path
distillation set-up can be used.
26 CHAPTER 3. VACUUM DISTILLATION

Perkin triangle distillation set-up

The Perkin triangle set-up (Image 5) uses a series of Teflon valves to allow the distilled fractions to be isolated from
the distillation flask without the main body of the distillation set-up being removed from either the vacuum or the
heat source, and thus can remain in a state of reflux.
To do this, the distillate receiver vessel is first isolated from the vacuum by means of the Teflon valves.
The vacuum over the sample is then replaced with an inert gas (such as nitrogen or argon) and the distillate receiver
can then be stoppered and removed from the system.

Vacuum distillation set-up using a short-path head

Vacuum distillation of moderately air/water-sensitive liquid can be done using standard Schlenk-line techniques (Im-
age 6). When assembling the set-up apparatus, all of the connecting lines are clamped so that they cannot pop off.
Once the apparatus is assembled, and the liquid to be distilled is in the still pot, the desired vacuum is established
in the system by using the vacuum connection on the short-path distillation head. Care is taken to prevent potential
“bumping” as the liquid in the still pot degases.
While establishing the vacuum, the flow of coolant is started through the short-path distillation head. Once the desired
vacuum is established, heat is applied to the still pot.
If needed, the first portion of distillate can be discarded by purging with inert gas and changing out the distillate
receiver.
When the distillation is complete: the heat is removed, the vacuum connection is closed, and inert gas is purged
through the distillation head and the distillate receiver. While under the inert gas purge, remove the distillate receiver
and cap it with an air-tight cap. The distillate receiver can be stored under vacuum or under inert gas by using the
side-arm on the distillation flask.

3.2 Industrial-scale applications


Industrial-scale vacuum distillation[8] has several advantages. Close boiling mixtures may require many equilibrium
stages to separate the key components. One tool to reduce the number of stages needed is to utilize vacuum
distillation.[9] Vacuum distillation columns (as depicted in Figures 2 and 3) typically used in oil refineries have di-
ameters ranging up to about 14 metres (46 feet), heights ranging up to about 50 metres (164 feet), and feed rates
ranging up to about 25,400 cubic metres per day (160,000 barrels per day).
Vacuum distillation increases the relative volatility of the key components in many applications. The higher the
relative volatility, the more separable are the two components; this connotes fewer stages in a distillation column in
order to effect the same separation between the overhead and bottoms products. Lower pressures increase relative
volatilities in most systems.
A second advantage of vacuum distillation is the reduced temperature requirement at lower pressures. For many
systems, the products degrade or polymerize at elevated temperatures.
Vacuum distillation can improve a separation by:

• Prevention of product degradation or polymer formation because of reduced pressure leading to lower tower
bottoms temperatures,

• Reduction of product degradation or polymer formation because of reduced mean residence time especially in
columns using packing rather than trays.

• Increasing capacity, yield, and purity.

Another advantage of vacuum distillation is the reduced capital cost, at the expense of slightly more operating cost.
Utilizing vacuum distillation can reduce the height and diameter, and thus the capital cost of a distillation column.
3.2. INDUSTRIAL-SCALE APPLICATIONS 27

Figure 2: Simplified animation of a typical dry vacuum distillation column as used in oil refineries

3.2.1 Vacuum distillation in petroleum refining

Petroleum crude oil is a complex mixture of hundreds of different hydrocarbon compounds generally having from 3
to 60 carbon atoms per molecule, although there may be small amounts of hydrocarbons outside that range.[10][11][12]
The refining of crude oil begins with distilling the incoming crude oil in a so-called atmospheric distillation column
operating at pressures slightly above atmospheric pressure.[8][10][11]
In distilling the crude oil, it is important not to subject the crude oil to temperatures above 370 to 380 °C because
the high molecular weight components in the crude oil will undergo thermal cracking and form petroleum coke at
temperatures above that. Formation of coke would result in plugging the tubes in the furnace that heats the feed stream
to the crude oil distillation column. Plugging would also occur in the piping from the furnace to the distillation column
as well as in the column itself.
The constraint imposed by limiting the column inlet crude oil to a temperature of less than 370 to 380 °C yields a
residual oil from the bottom of the atmospheric distillation column consisting entirely of hydrocarbons that boil above
370 to 380 °C.
To further distill the residual oil from the atmospheric distillation column, the distillation must be performed at
absolute pressures as low as 10 to 40 mmHg (also referred to as Torr) so as to limit the operating temperature to less
than 370 to 380 °C.
Figure 2 is a simplified process diagram of a petroleum refinery vacuum distillation column that depicts the internals
of the column and Figure 3 is a photograph of a large vacuum distillation column in a petroleum refinery.
The 10 to 40 mmHg absolute pressure in a vacuum distillation column increases the volume of vapor formed per
volume of liquid distilled. The result is that such columns have very large diameters.[13]
Distillation columns such those in Images 1 and 2, may have diameters of 15 meters or more, heights ranging up to
about 50 meters, and feed rates ranging up to about 25,400 cubic meters per day (160,000 barrels per day).
The vacuum distillation column internals must provide good vapor–liquid contacting while, at the same time, main-
taining a very low pressure increase from the top of the column top to the bottom. Therefore, the vacuum column
28 CHAPTER 3. VACUUM DISTILLATION

uses distillation trays only where withdrawing products from the side of the column (referred to as side draws). Most
of the column uses packing material for the vapor–liquid contacting because such packing has a lower pressure drop
than distillation trays. This packing material can be either structured sheet metal or randomly dumped packing such
as Raschig rings.
The absolute pressure of 10 to 40 mmHg in the vacuum column is most often achieved by using multiple stages of
steam jet ejectors.[14]
Many industries, other than the petroleum refining industry, use vacuum distillation on a much a smaller scale.

3.3 Molecular distillation


Molecular distillation is vacuum distillation below the pressure of 0.01 torr[15] (1.3 Pa). 0.01 torr is one order
of magnitude above high vacuum, where fluids are in the free molecular flow regime, i.e. the mean free path of
molecules is comparable to the size of the equipment. The gaseous phase no longer exerts significant pressure on the
substance to be evaporated, and consequently, rate of evaporation no longer depends on pressure. That is, because
the continuum assumptions of fluid dynamics no longer apply, mass transport is governed by molecular dynamics
rather than fluid dynamics. Thus, a short path between the hot surface and the cold surface is necessary, typically
by suspending a hot plate covered with a film of feed next to a cold plate with a line of sight in between. Molecular
distillation is used industrially for purification of oils.

3.4 Gallery
• A simple short path vacuum distillation apparatus

• Kugelrohr - a short path vacuum distillation apparatus

• Perkin triangle - for air-sensitive vacuum distillation

• Vacuum distillation apparatus

3.5 See also


• Continuous distillation

• Fractionating column

• Fractional distillation

• Kugelrohr

3.6 References
This article incorporates material from the Citizendium article "Vacuum distillation", which is licensed
under the Creative Commons Attribution-ShareAlike 3.0 Unported License but not under the GFDL.

[1] Laurence M. Harwood, Christopher J. Moody (13 June 1989). Experimental organic chemistry: Principles and Practice
(Illustrated edition ed.). WileyBlackwell. pp. 147–149. ISBN 978-0-632-02017-1.

[2] Distillation (CU Boulder Organic Chemistry Teaching Labs)

[3] Vacuum Distillation: New Method for Analyzing Organic Chemicals in a Wide Array of Samples (United States Environ-
mental Protection Agency)

[4] What is vacuum distillation? (Argonne National Laboratory's NEWTON Ask-A-Scientist)

[5] Operation of a Rotary Evaporator (Rotovap) (from the website of the University of British Columbia)
3.7. EXTERNAL LINKS 29

[6] SCAQMD Test method 302-91

[7] Energy Institute website page

[8] Kister, Henry Z. (1992). Distillation Design (1st Edition ed.). McGraw-Hill. ISBN 0-07-034909-6.

[9] Karl Kolmetz, Andrew W. Sloley et al. (2004), Designing Distillation Columns for Vacuum Service, 11th India Oil and Gas
Symposium and International Exhibition, September 2004, Mumbai, India (also published in Hydrocarbon Processing, May
2005)

[10] Gary, J.H. and Handwerk, G.E. (1984). Petroleum Refining Technology and Economics (2nd Edition ed.). Marcel Dekker,
Inc. ISBN 0-8247-7150-8.

[11] Leffler, W.L. (1985). Petroleum refining for the nontechnical person (2nd Edition ed.). PennWell Books. ISBN 0-87814-
280-0.

[12] James G, Speight (2006). The Chemistry and Technology of Petroleum (Fourth Edition ed.). CRC Press. 0-8493-9067-2.

[13] Karl Kolmetz, Andrew W. Sloley et al (2004), Designing Distillation Columns for Vacuum Service, 11th India Oil and Gas
Symposium and International Exhibition, September 2004, Mumbai, India (also published in Hydrocarbon Processing, May
2005)

[14] Photo gallery (from website of Graham Manufacturing Company)

[15] Vogel’s 5th ed.

3.7 External links


• D1160 Vacuum Distillation
30 CHAPTER 3. VACUUM DISTILLATION
Chapter 4

Hydrodesulfurization

Not to be confused with Flue-gas desulfurization.

Hydrodesulfurization (HDS) is a catalytic chemical process widely used to remove sulfur (S) from natural gas and
from refined petroleum products such as gasoline or petrol, jet fuel, kerosene, diesel fuel, and fuel oils.[1][2] The
purpose of removing the sulfur is to reduce the sulfur dioxide (SO
2) emissions that result from using those fuels in automotive vehicles, aircraft, railroad locomotives, ships, gas or oil
burning power plants, residential and industrial furnaces, and other forms of fuel combustion.
Another important reason for removing sulfur from the naphtha streams within a petroleum refinery is that sulfur, even
in extremely low concentrations, poisons the noble metal catalysts (platinum and rhenium) in the catalytic reforming
units that are subsequently used to upgrade the octane rating of the naphtha streams.
The industrial hydrodesulfurization processes include facilities for the capture and removal of the resulting hydrogen
sulfide (H
2S) gas. In petroleum refineries, the hydrogen sulfide gas is then subsequently converted into byproduct elemental
sulfur or sulfuric acid (H
2SO
4). In fact, the vast majority of the 64,000,000 metric tons of sulfur produced worldwide in 2005 was byproduct
sulfur from refineries and other hydrocarbon processing plants.[3][4]
An HDS unit in the petroleum refining industry is also often referred to as a hydrotreater.

4.1 History

Although some reactions involving catalytic hydrogenation of organic substances were already known, the property
of finely divided nickel to catalyze the fixation of hydrogen on hydrocarbon (ethylene, benzene) double bonds was
discovered by the French chemist Paul Sabatier in 1897.[5][6] Through this work, he found that unsaturated hydro-
carbons in the vapor phase could be converted into saturated hydrocarbons by using hydrogen and a catalytic metal,
laying the foundation of the modern catalytic hydrogenation process.
Soon after Sabatier’s work, a German chemist, Wilhelm Normann, found that catalytic hydrogenation could be used
to convert unsaturated fatty acids or glycerides in the liquid phase into saturated ones. He was awarded a patent in
Germany in 1902[7] and in Britain in 1903,[8] which was the beginning of what is now a worldwide industry.
In the mid-1950s, the first noble metal catalytic reforming process (the Platformer process) was commercialized. At
the same time, the catalytic hydrodesulfurization of the naphtha feed to such reformers was also commercialized. In
the decades that followed, various proprietary catalytic hydrodesulfurization processes, such as the one depicted in
the flow diagram below, have been commercialized. Currently, virtually all of the petroleum refineries worldwide
have one or more HDS units.
By 2006, miniature microfluidic HDS units had been implemented for treating JP-8 jet fuel to produce clean feed
stock for a fuel cell hydrogen reformer.[9] By 2007, this had been integrated into an operating 5 kW fuel cell generation
system.[10]

31
32 CHAPTER 4. HYDRODESULFURIZATION

4.2 Process chemistry


Hydrogenation is a class of chemical reactions in which the net result is the addition of hydrogen (H). Hydrogenolysis
is a type of hydrogenation and results in the cleavage of the C-X chemical bond, where C is a carbon atom and X is a
sulfur (S), nitrogen (N) or oxygen (O) atom. The net result of a hydrogenolysis reaction is the formation of C-H and
H-X chemical bonds. Thus, hydrodesulfurization is a hydrogenolysis reaction. Using ethanethiol (C
2H
5SH), a sulfur compound present in some petroleum products, as an example, the hydrodesulfurization reaction can
be simply expressed as

For the mechanistic aspects of, and the catalysts used in this reaction see the section catalysts and mechanisms

4.3 Process description


In an industrial hydrodesulfurization unit, such as in a refinery, the hydrodesulfurization reaction takes place in a
fixed-bed reactor at elevated temperatures ranging from 300 to 400 °C and elevated pressures ranging from 30 to
130 atmospheres of absolute pressure, typically in the presence of a catalyst consisting of an alumina base impreg-
nated with cobalt and molybdenum (usually called a CoMo catalyst). Occasionally, a combination of nickel and
molybdenum (called NiMo) is used, in addition to the CoMo catalyst, for specific difficult-to-treat feed stocks, such
as those containing a high level of chemically bound nitrogen.
The image below is a schematic depiction of the equipment and the process flow streams in a typical refinery HDS
unit.

Schematic diagram of a typical Hydrodesulfurization (HDS) unit in a petroleum refinery

The liquid feed (at the bottom left in the diagram) is pumped up to the required elevated pressure and is joined
by a stream of hydrogen-rich recycle gas. The resulting liquid-gas mixture is preheated by flowing through a heat
exchanger. The preheated feed then flows through a fired heater where the feed mixture is totally vaporized and
heated to the required elevated temperature before entering the reactor and flowing through a fixed-bed of catalyst
where the hydrodesulfurization reaction takes place.
4.4. SULFUR COMPOUNDS IN REFINERY HDS FEEDSTOCKS 33

The hot reaction products are partially cooled by flowing through the heat exchanger where the reactor feed was
preheated and then flows through a water-cooled heat exchanger before it flows through the pressure controller (PC)
and undergoes a pressure reduction down to about 3 to 5 atmospheres. The resulting mixture of liquid and gas enters
the gas separator vessel at about 35 °C and 3 to 5 atmospheres of absolute pressure.
Most of the hydrogen-rich gas from the gas separator vessel is recycle gas, which is routed through an amine contactor
for removal of the reaction product H
2S that it contains. The H
2S-free hydrogen-rich gas is then recycled back for reuse in the reactor section. Any excess gas from the gas separator
vessel joins the sour gas from the stripping of the reaction product liquid.
The liquid from the gas separator vessel is routed through a reboiled stripper distillation tower. The bottoms product
from the stripper is the final desulfurized liquid product from hydrodesulfurization unit.
The overhead sour gas from the stripper contains hydrogen, methane, ethane, hydrogen sulfide, propane, and, perhaps,
some butane and heavier components. That sour gas is sent to the refinery’s central gas processing plant for removal
of the hydrogen sulfide in the refinery’s main amine gas treating unit and through a series of distillation towers for
recovery of propane, butane and pentane or heavier components. The residual hydrogen, methane, ethane, and some
propane is used as refinery fuel gas. The hydrogen sulfide removed and recovered by the amine gas treating unit is
subsequently converted to elemental sulfur in a Claus process unit or to sulfuric acid in a wet sulfuric acid process or
in the conventional Contact Process.
Note that the above description assumes that the HDS unit feed contains no olefins. If the feed does contain olefins
(for example, the feed is a naphtha derived from a refinery fluid catalytic cracker (FCC) unit), then the overhead gas
from the HDS stripper may also contain some ethene, propene, butenes and pentenes, or heavier components.
It should also be noted that the amine solution to and from the recycle gas contactor comes from and is returned to
the refinery’s main amine gas treating unit.

4.4 Sulfur compounds in refinery HDS feedstocks


The refinery HDS feedstocks (naphtha, kerosene, diesel oil, and heavier oils) contain a wide range of organic sulfur
compounds, including thiols, thiophenes, organic sulfides and disulfides, and many others. These organic sulfur
compounds are products of the degradation of sulfur containing biological components, present during the natural
formation of the fossil fuel, petroleum crude oil.
When the HDS process is used to desulfurize a refinery naphtha, it is necessary to remove the total sulfur down to
the parts per million range or lower in order to prevent poisoning the noble metal catalysts in the subsequent catalytic
reforming of the naphthas.
When the process is used for desulfurizing diesel oils, the latest environmental regulations in the United States and
Europe, requiring what is referred to as ultra-low sulfur diesel (ULSD), in turn requires that very deep hydrodesulfu-
rization is needed. In the very early 2000s, the governmental regulatory limits for highway vehicle diesel was within
the range of 300 to 500 ppm by weight of total sulfur. As of 2006, the total sulfur limit for highway diesel is in the
range of 15 to 30 ppm by weight.[11]

4.4.1 Thiophenes
A family of substrates that are particularly common in petroleum are the aromatic sulfur-containing heterocycles
called thiophenes. Many kinds of thiophenes occur in petroleum ranging from thiophene itself to more condensed
derivatives called benzothiophenes and dibenzothiophenes. Thiophene itself and its alkyl derivatives are easier to
hydrogenolyse, whereas dibenzothiophene, especially its 4,6-disubstituted derivatives, are considered the most chal-
lenging substrates. Benzothiophenes are midway between the simple thiophenes and dibenzothiophenes in their
susceptibility to HDS.

4.5 Catalysts and mechanisms


The main HDS catalysts are based on molybdenum disulfide (MoS
2) together with smaller amounts of other metals.[12] The nature of the sites of catalytic activity remains an active
34 CHAPTER 4. HYDRODESULFURIZATION

area of investigation, but it is generally assumed basal planes of the MoS


2 structure are not relevant to catalysis, rather the edges or rims of these sheet.[13] At the edges of the MoS
2 crystallites, the molybdenum centre can stabilize a coordinatively unsaturated site (CUS), also known as an anion
vacancy. Substrates, such as thiophene, bind to this site and undergo a series a reactions that result in both C-S
scission and C=C hydrogenation. Thus, the hydrogen serves multiple roles—generation of anion vacancy by removal
of sulfide, hydrogenation, and hydrogenolysis. A simplified diagram for the cycle is shown:

Simplified diagram of a HDS cycle for thiophene

4.5.1 Catalysts
Most metals catalyse HDS, but it is those at the middle of the transition metal series that are most active. Ruthenium
disulfide appears to be the single most active catalyst, but binary combinations of cobalt and molybdenum are also
highly active.[14] Aside from the basic cobalt-modified MoS2 catalyst, nickel and tungsten are also used, depending
on the nature of the feed. For example, Ni-W catalysts are more effective for hydrodenitrogenation.

4.5.2 Supports
Metal sulfides are “supported” on materials with high surface areas. A typical support for HDS catalyst is γ-alumina.
The support allows the more expensive catalyst to be more widely distributed, giving rise to a larger fraction of the
MoS
2 that is catalytically active. The interaction between the support and the catalyst is an area of intense interest, since
the support is often not fully inert but participates in the catalysis.

4.6 Other uses


The basic hydrogenolysis reaction has a number of uses other than hydrodesulfurization.

4.6.1 Hydrodenitrogenation
The hydrogenolysis reaction is also used to reduce the nitrogen content of a petroleum stream in a process referred
to as hydrodenitrogenation (HDN). The process flow is the same as that for an HDS unit.
Using pyridine (C
5H
5N), a nitrogen compound present in some petroleum fractionation products, as an example, the hydrodenitrogenation
reaction has been postulated as occurring in three steps:[15][16]
4.7. SEE ALSO 35

and the overall reaction may be simply expressed as:

Many HDS units for desulfurizing naphthas within petroleum refineries are actually simultaneously denitrogenating
to some extent as well.

4.6.2 Saturation of olefins


The hydrogenolysis reaction may also be used to saturate or convert olefins (alkenes) into paraffins (alkanes). The
process used is the same as for an HDS unit.
As an example, the saturation of the olefin pentene can be simply expressed as:

Some hydrogenolysis units within a petroleum refinery or a petrochemical plant may be used solely for the saturation
of olefins or they may be used for simultaneously desulfurizing as well as denitrogenating and saturating olefins to
some extent.

4.6.3 Hydrogenation in the food industry


Further information: Hydrogenation, Wilhelm Normann and Trans fat

The food industry uses hydrogenation to completely or partially saturate the unsaturated fatty acids in liquid vegetable
fats and oils to convert them into solid or semi-solid fats, such as those in margarine and shortening.

4.7 See also


• Claus process
• Hydrogen pinch
• Timeline of hydrogen technologies

4.8 References
[1] Gary, J.H. and Handwerk, G.E. (1984). Petroleum Refining Technology and Economics (2nd ed.). Marcel Dekker, Inc.
ISBN 0-8247-7150-8.

[2] Hydrodesulfurization Technologies and Costs Nancy Yamaguchi, Trans Energy Associates, William and Flora Hewlett
Foundation Sulfur Workshop, Mexico City, May 29–30, 2003

[3] Sulfur production report by the United States Geological Survey

[4] Discussion of recovered byproduct sulfur

[5] C.R.Acad.Sci. 1897, 132, 210

[6] C.R.Acad.Sci. 1901, 132, 210

[7] DE Patent DE141029 (Espacenet, record not available)

[8] UK Patent GB190301515 GB190301515 (Espacenet)

[9] Microchannel HDS (March 2006)


36 CHAPTER 4. HYDRODESULFURIZATION

[10] Fuel cells help make noisy, hot generators a thing of the past (December 2007) Pacific Northwest National Laboratory

[11] Diesel Sulfur published online by the National Petrochemical & Refiners Association (NPRA)

[12] Topsøe, H.; Clausen, B. S.; Massoth, F. E., Hydrotreating Catalysis, Science and Technology, Springer-Verlag: Berlin,
1996.

[13] Daage, M.; Chianelli, R. R., “Structure-Function Relations in Molybdenum Sulfide Catalysts - the Rim-Edge Model”, J. of
Catalysis, 1994, 149, 414-427.

[14] Chianelli, R. R.; Berhault, G.; Raybaud, P.; Kasztelan, S.; Hafner, J. and Toulhoat, H., “Periodic trends in hydrodesulfur-
ization: in support of the Sabatier principle”, Applied Catalysis, A, 2002, volume 227, pages 83-96.

[15] Kinetics and Interactions of the Simultaneous Catalytic Hydrodenitrogenation of Pyridine and Hydrodesulfurization of Thio-
phene (John Wilkins, PhD Thesis, MIT, 1977)

[16] Simultaneous Catalytic Hydrodenitrogenation of Pyridine and Hydrodesulfurization of Thiophene (Satterfield,C.N., Modell,
M. and Wilkens, J.A., Ind. Eng. Chem. Process Des. Dev., 1980 Vol. 19, pages 154-160)

4.9 External links


• Criterion Catalysts (Hydroprocessing Catalyst Supplier)
• Haldor Topsoe (Catalyzing Your Business)

• Albemarle Catalyst Company (Petrochemical catalysts supplier)


• UOP-Honeywell (Engineering design and construction of large-scale, industrial HDS plants)

• Hydrogenation for Low Trans and High Conjugated Fatty Acids by E.S. Jang, M.Y. Jung, D.B. Min, Compre-
hensive Reviews in Food Science and Food Safety, Vol.1, 2005

• Oxo Alcohols (Engineered and constructed by Aker Kvaerner)


• Catalysts and technology for Oxo-Alcohols
Chapter 5

Catalytic reforming

A catalytic reformer unit in a petroleum refinery. © BP p.l.c

Catalytic reforming is a chemical process used to convert petroleum refinery naphthas distilled from crude oil (typ-
ically having low octane ratings) into high-octane liquid products called reformates, which are premium blending
stocks for high-octane gasoline. The process converts low-octane linear hydrocarbons (paraffins) into branched alka-

37
38 CHAPTER 5. CATALYTIC REFORMING

nes (isoparaffins) and cyclic naphthenes, which are then partially dehydrogenated to produce high-octane aromatic
hydrocarbons. The dehydrogenation also produces significant amounts of byproduct hydrogen gas, which is fed into
other refinery processes such as hydrocracking. A side reaction is hydrogenolysis, which produces light hydrocarbons
of lower value, such as methane, ethane, propane and butanes.
In addition to a gasoline blending stock, reformate is the main source of aromatic bulk chemicals such as benzene,
toluene, xylene and ethylbenzene which have diverse uses, most importantly as raw materials for conversion into
plastics. However, the benzene content of reformate makes it carcinogenic, which has led to governmental regulations
effectively requiring further processing to reduce its benzene content.
This process is quite different from and not to be confused with the catalytic steam reforming process used industrially
to produce products such as hydrogen, ammonia, and methanol from natural gas, naphtha or other petroleum-derived
feedstocks. Nor is this process to be confused with various other catalytic reforming processes that use methanol or
biomass-derived feedstocks to produce hydrogen for fuel cells or other uses.

5.1 History
In the 1940s, Vladimir Haensel,[1] a research chemist working for Universal Oil Products (UOP), developed a catalytic
reforming process using a catalyst containing platinum. Haensel’s process was subsequently commercialized by UOP
in 1949 for producing a high octane gasoline from low octane naphthas and the UOP process become known as the
Platforming process.[2] The first Platforming unit was built in 1949 at the refinery of the Old Dutch Refining Company
in Muskegon, Michigan.
In the years since then, many other versions of the process have been developed by some of the major oil companies
and other organizations. Today, the large majority of gasoline produced worldwide is derived from the catalytic
reforming process.
To name a few of the other catalytic reforming versions that were developed, all of which utilized a platinum and/or
a rhenium catalyst:

• Rheniforming: Developed by Chevron Oil Company.

• Powerforming: Developed by Esso Oil Company, currently known as ExxonMobil.

• Magnaforming: Developed by Engelhard and Atlantic Richfield Oil Company.

• Ultraforming: Developed by Standard Oil of Indiana, now a part of the British Petroleum Company.

• Houdriforming: Developed by the Houdry Process Corporation.

• CCR Platforming: A Platforming version, designed for continuous catalyst regeneration, developed by UOP.

• Octanizing: A catalytic reforming version developed by Axens, a subsidiary of Institut francais du petrole (IFP),
designed for continuous catalyst regeneration.

5.2 Chemistry
Before describing the reaction chemistry of the catalytic reforming process as used in petroleum refineries, the typical
naphthas used as catalytic reforming feedstocks will be discussed.

5.2.1 Typical naphtha feedstocks


A petroleum refinery includes many unit operations and unit processes. The first unit operation in a refinery is the
continuous distillation of the petroleum crude oil being refined. The overhead liquid distillate is called naphtha and
will become a major component of the refinery’s gasoline (petrol) product after it is further processed through a
catalytic hydrodesulfurizer to remove sulfur-containing hydrocarbons and a catalytic reformer to reform its hydro-
carbon molecules into more complex molecules with a higher octane rating value. The naphtha is a mixture of very
many different hydrocarbon compounds. It has an initial boiling point of about 35 °C and a final boiling point of
5.2. CHEMISTRY 39

about 200 °C, and it contains paraffin, naphthene (cyclic paraffins) and aromatic hydrocarbons ranging from those
containing 4 carbon atoms to those containing about 10 or 11 carbon atoms.
The naphtha from the crude oil distillation is often further distilled to produce a “light” naphtha containing most (but
not all) of the hydrocarbons with 6 or fewer carbon atoms and a “heavy” naphtha containing most (but not all) of the
hydrocarbons with more than 6 carbon atoms. The heavy naphtha has an initial boiling point of about 140 to 150 °C
and a final boiling point of about 190 to 205 °C. The naphthas derived from the distillation of crude oils are referred
to as “straight-run” naphthas.
It is the straight-run heavy naphtha that is usually processed in a catalytic reformer because the light naphtha has
molecules with 6 or fewer carbon atoms which, when reformed, tend to crack into butane and lower molecular weight
hydrocarbons which are not useful as high-octane gasoline blending components. Also, the molecules with 6 carbon
atoms tend to form aromatics which is undesirable because governmental environmental regulations in a number of
countries limit the amount of aromatics (most particularly benzene) that gasoline may contain.[3][4][5]
It should be noted that there are a great many petroleum crude oil sources worldwide and each crude oil has its own
unique composition or “assay”. Also, not all refineries process the same crude oils and each refinery produces its own
straight-run naphthas with their own unique initial and final boiling points. In other words, naphtha is a generic term
rather than a specific term.
The table just below lists some fairly typical straight-run heavy naphtha feedstocks, available for catalytic reforming,
derived from various crude oils. It can be seen that they differ significantly in their content of paraffins, naphthenes
and aromatics:
.....
Some refinery naphthas include olefinic hydrocarbons, such as naphthas derived from the fluid catalytic cracking
and coking processes used in many refineries. Some refineries may also desulfurize and catalytically reform those
naphthas. However, for the most part, catalytic reforming is mainly used on the straight-run heavy naphthas, such as
those in the above table, derived from the distillation of crude oils.

5.2.2 The reaction chemistry


There are many chemical reactions that occur in the catalytic reforming process, all of which occur in the presence
of a catalyst and a high partial pressure of hydrogen. Depending upon the type or version of catalytic reforming used
as well as the desired reaction severity, the reaction conditions range from temperatures of about 495 to 525 °C and
from pressures of about 5 to 45 atm.[10]
The commonly used catalytic reforming catalysts contain noble metals such as platinum and/or rhenium, which are
very susceptible to poisoning by sulfur and nitrogen compounds. Therefore, the naphtha feedstock to a catalytic
reformer is always pre-processed in a hydrodesulfurization unit which removes both the sulfur and the nitrogen com-
pounds. Most catalysts require both sulphur and nitrogen content to be lower than 1 ppm.
The four major catalytic reforming reactions are:[11]

1: The dehydrogenation of naphthenes to convert them into aromatics as exemplified in the conversion
methylcyclohexane (a naphthene) to toluene (an aromatic), as shown below:

+ 3 H2

2: The isomerization of normal paraffins to isoparaffins as exemplified in the conversion of normal octane
to 2,5-Dimethylhexane (an isoparaffin), as shown below:

3: The dehydrogenation and aromatization of paraffins to aromatics (commonly called dehydrocycliza-


tion) as exemplified in the conversion of normal heptane to toluene, as shown below:
40 CHAPTER 5. CATALYTIC REFORMING

4: The hydrocracking of paraffins into smaller molecules as exemplified by the cracking of normal
heptane into isopentane and ethane, as shown below:

The hydrocracking of paraffins is the only one of the above four major reforming reactions that consumes hydrogen.
The isomerization of normal paraffins does not consume or produce hydrogen. However, both the dehydrogenation
of naphthenes and the dehydrocyclization of paraffins produce hydrogen. The overall net production of hydrogen
in the catalytic reforming of petroleum naphthas ranges from about 50 to 200 cubic meters of hydrogen gas (at 0
°C and 1 atm) per cubic meter of liquid naphtha feedstock. In the United States customary units, that is equivalent
to 300 to 1200 cubic feet of hydrogen gas (at 60 °F and 1 atm) per barrel of liquid naphtha feedstock.[12] In many
petroleum refineries, the net hydrogen produced in catalytic reforming supplies a significant part of the hydrogen used
elsewhere in the refinery (for example, in hydrodesulfurization processes). The hydrogen is also necessary in order
to hydrogenolyze any polymers that form on the catalyst.
In practice, the higher the content of naphtenes in the naphtha feedstock, the better will be the quality of the reformate
and the higher the production of hydrogen. Crude oils containing the best naphtha for reforming are typically from
Western Africa or the North Sea, such as Bonny light or Troll.

5.3 Process description


The most commonly used type of catalytic reforming unit has three reactors, each with a fixed bed of catalyst, and all
of the catalyst is regenerated in situ during routine catalyst regeneration shutdowns which occur approximately once
each 6 to 24 months. Such a unit is referred to as a semi-regenerative catalytic reformer (SRR).
Some catalytic reforming units have an extra spare or swing reactor and each reactor can be individually isolated
so that any one reactor can be undergoing in situ regeneration while the other reactors are in operation. When that
reactor is regenerated, it replaces another reactor which, in turn, is isolated so that it can then be regenerated. Such
units, referred to as cyclic catalytic reformers, are not very common. Cyclic catalytic reformers serve to extend the
period between required shutdowns.
5.3. PROCESS DESCRIPTION 41

The latest and most modern type of catalytic reformers are called continuous catalyst regeneration (CCR) reformers.
Such units are characterized by continuous in-situ regeneration of part of the catalyst in a special regenerator, and by
continuous addition of the regenerated catalyst to the operating reactors. As of 2006, two CCR versions available:
UOP’s CCR Platformer process[13] and Axens’ Octanizing process.[14] The installation and use of CCR units is rapidly
increasing.
Many of the earliest catalytic reforming units (in the 1950s and 1960s) were non-regenerative in that they did not
perform in situ catalyst regeneration. Instead, when needed, the aged catalyst was replaced by fresh catalyst and the
aged catalyst was shipped to catalyst manufacturers to be either regenerated or to recover the platinum content of the
aged catalyst. Very few, if any, catalytic reformers currently in operation are non-regenerative.
The process flow diagram below depicts a typical semi-regenerative catalytic reforming unit.

Schematic diagram of a typical semi-regenerative catalytic reformer unit in a petroleum refinery

The liquid feed (at the bottom left in the diagram) is pumped up to the reaction pressure (5–45 atm) and is joined
by a stream of hydrogen-rich recycle gas. The resulting liquid–gas mixture is preheated by flowing through a heat
exchanger. The preheated feed mixture is then totally vaporized and heated to the reaction temperature (495–520
°C) before the vaporized reactants enter the first reactor. As the vaporized reactants flow through the fixed bed of
catalyst in the reactor, the major reaction is the dehydrogenation of naphthenes to aromatics (as described earlier
herein) which is highly endothermic and results in a large temperature decrease between the inlet and outlet of the
reactor. To maintain the required reaction temperature and the rate of reaction, the vaporized stream is reheated in
the second fired heater before it flows through the second reactor. The temperature again decreases across the second
reactor and the vaporized stream must again be reheated in the third fired heater before it flows through the third
reactor. As the vaporized stream proceeds through the three reactors, the reaction rates decrease and the reactors
therefore become larger. At the same time, the amount of reheat required between the reactors becomes smaller.
Usually, three reactors are all that is required to provide the desired performance of the catalytic reforming unit.
Some installations use three separate fired heaters as shown in the schematic diagram and some installations use a
single fired heater with three separate heating coils.
The hot reaction products from the third reactor are partially cooled by flowing through the heat exchanger where the
feed to the first reactor is preheated and then flow through a water-cooled heat exchanger before flowing through the
pressure controller (PC) into the gas separator.
Most of the hydrogen-rich gas from the gas separator vessel returns to the suction of the recycle hydrogen gas com-
pressor and the net production of hydrogen-rich gas from the reforming reactions is exported for use in the other
refinery processes that consume hydrogen (such as hydrodesulfurization units and/or a hydrocracker unit).
The liquid from the gas separator vessel is routed into a fractionating column commonly called a stabilizer. The over-
head offgas product from the stabilizer contains the byproduct methane, ethane, propane and butane gases produced
by the hydrocracking reactions as explained in the above discussion of the reaction chemistry of a catalytic reformer,
42 CHAPTER 5. CATALYTIC REFORMING

and it may also contain some small amount of hydrogen. That offgas is routed to the refinery’s central gas processing
plant for removal and recovery of propane and butane. The residual gas after such processing becomes part of the
refinery’s fuel gas system.
The bottoms product from the stabilizer is the high-octane liquid reformate that will become a component of the
refinery’s product gasoline. Reformate can be blended directly in the gasoline pool but often it is separated in two
or more streams. A common refining scheme consists in fractionating the reformate in two streams, light and heavy
reformate. The light reformate has lower octane and can be used as isomerization feedstock if this unit is available.
The heavy reformate is high in octane and low in benzene, hence it is an excellent blending component for the gasoline
pool.
Benzene is often removed with a specific operation to reduce the content of benzene in the reformate as the finished
gasoline has often an upper limit of benzene content (in the UE this is 1% volume). The benzene extracted can be
marketed as feedstock for the chemical industry.

5.4 Catalysts and mechanisms


Most catalytic reforming catalysts contain platinum or rhenium on a silica or silica-alumina support base, and some
contain both platinum and rhenium. Fresh catalyst is chlorided (chlorinated) prior to use.
The noble metals (platinum and rhenium) are considered to be catalytic sites for the dehydrogenation reactions and
the chlorinated alumina provides the acid sites needed for isomerization, cyclization and hydrocracking reactions.[11]
The biggest care has to be exercised during the chlorination. Indeed if not chlorinated (or insufficiently chlorinated)
the platinum and rhenium in the catalyst would be reduced almost immediately to metallic state by the hydrogen in
the vapour phase. On the other an excessive chlorination could depress excessively the activity of the catalyst.
The activity (i.e., effectiveness) of the catalyst in a semi-regenerative catalytic reformer is reduced over time during
operation by carbonaceous coke deposition and chloride loss. The activity of the catalyst can be periodically regener-
ated or restored by in situ high temperature oxidation of the coke followed by chlorination. As stated earlier herein,
semi-regenerative catalytic reformers are regenerated about once per 6 to 24 months. The higher the severity of the
reacting conditions (temperature), the higher is the octane of the produced reformate but also the shorter will be the
duration of the cycle between two regenerations. Catalyst’s cycle duration is also very dependent on the quality of
the feedstock. However, independently of the crude oil used in the refinery, all catalysts require a maximum final
boiling point of the naphtha feedstock of 180 °C.
Normally, the catalyst can be regenerated perhaps 3 or 4 times before it must be returned to the manufacturer for
reclamation of the valuable platinum and/or rhenium content.[11]

5.5 Economics

5.6 References
[1] A Biographical Memoir of Vladimir Haensel written by Stanley Gembiki, published by the National Academy of Sciences
in 2006.

[2] Platforming described on UOP’s website

[3] Canadian regulations on benzene in gasoline

[4] United Kingdom regulations on benzene in gasoline

[5] USA regulations on benzene in gasoline

[6] Barrow Island crude oil assay

[7] Mutineer-Exeter crude oil assay

[8] CPC Blend crude oil assay

[9] Draugen crude oil assay


5.7. EXTERNAL LINKS 43

[10] OSHA Technical Manual, Section IV, Chapter 2, Petroleum refining Processes (A publication of the Occupational Safety
and Health Administration)

[11] Gary, J.H. and Handwerk, G.E. (1984). Petroleum Refining Technology and Economics (2nd Edition ed.). Marcel Dekker,
Inc. ISBN 0-8247-7150-8.

[12] US Patent 5011805, Dehydrogenation, dehydrocyclization and reforming catalyst (Inventor: Ralph Dessau, Assignee: Mobil
Oil Corporation)

[13] CCR Platforming (UOP website)

[14] Octanizing Options (Axens website)

5.7 External links


• Oil Refinery Processes, A Brief Overview

• Colorado School of Mines, Lecture Notes (Chapter 10, Refining Processes, Catalytic Refinery by John Jechura,
Adjunct Professor)

• Students’ Guide to Refining (scroll down to Platforming)


• Modern Refinery Website of Delft University of Technology, Netherlands (use search function for Reforming)

• Major scientific and technical challenges about development of new refining processes (IFP website)
Chapter 6

Fluid catalytic cracking

Fluid catalytic cracking (FCC) is one of the most important conversion processes used in petroleum refineries. It
is widely used to convert the high-boiling, high-molecular weight hydrocarbon fractions of petroleum crude oils to
more valuable gasoline, olefinic gases, and other products.[1][2][3] Cracking of petroleum hydrocarbons was originally
done by thermal cracking, which has been almost completely replaced by catalytic cracking because it produces more
gasoline with a higher octane rating. It also produces byproduct gases that are more olefinic, and hence more valuable,
than those produced by thermal cracking.
The feedstock to an FCC is usually that portion of the crude oil that has an initial boiling point of 340 °C or higher at
atmospheric pressure and an average molecular weight ranging from about 200 to 600 or higher. This portion of crude
oil is often referred to as heavy gas oil or vacuum gas oil (HVGO). The FCC process vaporizes and breaks the long-
chain molecules of the high-boiling hydrocarbon liquids into much shorter molecules by contacting the feedstock, at
high temperature and moderate pressure, with a fluidized powdered catalyst.
In effect, refineries use fluid catalytic cracking to correct the imbalance between the market demand for gasoline and
the excess of heavy, high boiling range products resulting from the distillation of crude oil.
As of 2006, FCC units were in operation at 400 petroleum refineries worldwide and about one-third of the crude oil
refined in those refineries is processed in an FCC to produce high-octane gasoline and fuel oils.[2][4] During 2007,
the FCC units in the United States processed a total of 5,300,000 barrels (840,000 m3 ) per day of feedstock[5] and
FCC units worldwide processed about twice that amount.

6.1 Flow diagram and process description


The modern FCC units are all continuous processes which operate 24 hours a day for as long as 3 to 5 years between
scheduled shutdowns for routine maintenance.
There are several different proprietary designs that have been developed for modern FCC units. Each design is
available under a license that must be purchased from the design developer by any petroleum refining company desiring
to construct and operate an FCC of a given design.
There are two different configurations for an FCC unit: the “stacked” type where the reactor and the catalyst re-
generator are contained in a single vessel with the reactor above the catalyst regenerator and the “side-by-side” type
where the reactor and catalyst regenerator are in two separate vessels. These are the major FCC designers and
licensors:[1][3][4][6]
Side-by-side configuration:

• CB&I
• ExxonMobil Research and Engineering (EMRE)
• Shell Global Solutions
• Axens / Stone & Webster Process Technology — currently owned by Technip
• Universal Oil Products (UOP) — currently fully owned subsidiary of Honeywell

44
6.1. FLOW DIAGRAM AND PROCESS DESCRIPTION 45

A typical fluid catalytic cracking unit in a petroleum refinery.

Stacked configuration:

• Kellogg Brown & Root (KBR)

Each of the proprietary design licensors claims to have unique features and advantages. A complete discussion of the
relative advantages of each of the processes is beyond the scope of this article. Suffice it to say that all of the licensors
have designed and constructed FCC units that have operated quite satisfactorily.
46 CHAPTER 6. FLUID CATALYTIC CRACKING

6.1.1 Reactor and Regenerator

The reactor and regenerator are considered to be the heart of the fluid catalytic cracking unit. The schematic flow
diagram of a typical modern FCC unit in Figure 1 below is based upon the “side-by-side” configuration. The pre-
heated high-boiling petroleum feedstock (at about 315 to 430 °C) consisting of long-chain hydrocarbon molecules is
combined with recycle slurry oil from the bottom of the distillation column and injected into the catalyst riser where
it is vaporized and cracked into smaller molecules of vapor by contact and mixing with the very hot powdered catalyst
from the regenerator. All of the cracking reactions take place in the catalyst riser within a period of 2–4 seconds.
The hydrocarbon vapors “fluidize” the powdered catalyst and the mixture of hydrocarbon vapors and catalyst flows
upward to enter the reactor at a temperature of about 535 °C and a pressure of about 1.72 barg.
The reactor is a vessel in which the cracked product vapors are: (a) separated from the so-called spent catalyst by
flowing through a set of two-stage cyclones within the reactor and (b) the spent catalyst flows downward through a
steam stripping section to remove any hydrocarbon vapors before the spent catalyst returns to the catalyst regenerator.
The flow of spent catalyst to the regenerator is regulated by a slide valve in the spent catalyst line.
Since the cracking reactions produce some carbonaceous material (referred to as catalyst coke) that deposits on the
catalyst and very quickly reduces the catalyst reactivity, the catalyst is regenerated by burning off the deposited coke
with air blown into the regenerator. The regenerator operates at a temperature of about 715 °C and a pressure of
about 2.41 barg. The combustion of the coke is exothermic and it produces a large amount of heat that is partially
absorbed by the regenerated catalyst and provides the heat required for the vaporization of the feedstock and the
endothermic cracking reactions that take place in the catalyst riser. For that reason, FCC units are often referred to
as being 'heat balanced'.
The hot catalyst (at about 715 °C) leaving the regenerator flows into a catalyst withdrawal well where any entrained
combustion flue gases are allowed to escape and flow back into the upper part to the regenerator. The flow of regen-
erated catalyst to the feedstock injection point below the catalyst riser is regulated by a slide valve in the regenerated
catalyst line. The hot flue gas exits the regenerator after passing through multiple sets of two-stage cyclones that
remove entrained catalyst from the flue gas,
The amount of catalyst circulating between the regenerator and the reactor amounts to about 5 kg per kg of feedstock,
which is equivalent to about 4.66 kg per litre of feedstock.[1][7] Thus, an FCC unit processing 75,000 barrels per day
(11,900 m3 /d) will circulate about 55,900 tonnes per day of catalyst.

6.1.2 Distillation column

The reaction product vapors (at 535 °C and a pressure of 1.72 barg) flow from the top of the reactor to the bottom
section of the distillation column (commonly referred to as the main fractionator) where they are distilled into the
FCC end products of cracked naphtha, fuel oil, and offgas. After further processing for removal of sulfur compounds,
the cracked naphtha becomes a high-octane component of the refinery’s blended gasolines.
The main fractionator offgas is sent to what is called a gas recovery unit where it is separated into butanes and
butylenes, propane and propylene, and lower molecular weight gases (hydrogen, methane, ethylene and ethane). Some
FCC gas recovery units may also separate out some of the ethane and ethylene.
Although the schematic flow diagram above depicts the main fractionator as having only one sidecut stripper and one
fuel oil product, many FCC main fractionators have two sidecut strippers and produce a light fuel oil and a heavy
fuel oil. Likewise, many FCC main fractionators produce a light cracked naphtha and a heavy cracked naphtha. The
terminology light and heavy in this context refers to the product boiling ranges, with light products having a lower
boiling range than heavy products.
The bottom product oil from the main fractionator contains residual catalyst particles which were not completely
removed by the cyclones in the top of the reactor. For that reason, the bottom product oil is referred to as a slurry
oil. Part of that slurry oil is recycled back into the main fractionator above the entry point of the hot reaction product
vapors so as to cool and partially condense the reaction product vapors as they enter the main fractionator. The
remainder of the slurry oil is pumped through a slurry settler. The bottom oil from the slurry settler contains most
of the slurry oil catalyst particles and is recycled back into the catalyst riser by combining it with the FCC feedstock
oil. The so-called clarified slurry oil or decant oil is withdrawn from the top of slurry settler for use elsewhere in the
refinery, as a heavy fuel oil blending component, or as carbon black feedstock.
6.1. FLOW DIAGRAM AND PROCESS DESCRIPTION 47

Figure 1: A schematic flow diagram of a Fluid Catalytic Cracking unit as used in petroleum refineries

6.1.3 Regenerator flue gas

Depending on the choice of FCC design, the combustion in the regenerator of the coke on the spent catalyst may or
may not be complete combustion to carbon dioxide CO
2. The combustion air flow is controlled so as to provide the desired ratio of carbon monoxide (CO) to carbon dioxide
for each specific FCC design.[1][4]
In the design shown in Figure 1, the coke has only been partially combusted to CO
2. The combustion flue gas (containing CO and CO
2) at 715 °C and at a pressure of 2.41 barg is routed through a secondary catalyst separator containing swirl tubes
designed to remove 70 to 90 percent of the particulates in the flue gas leaving the regenerator.[8] This is required to
prevent erosion damage to the blades in the turbo-expander that the flue gas is next routed through.
The expansion of flue gas through a turbo-expander provides sufficient power to drive the regenerator’s combustion
air compressor. The electrical motor-generator can consume or produce electrical power. If the expansion of the
flue gas does not provide enough power to drive the air compressor, the electric motor/generator provides the needed
additional power. If the flue gas expansion provides more power than needed to drive the air compressor, than
the electric motor/generator converts the excess power into electric power and exports it to the refinery’s electrical
system.[3]
The expanded flue gas is then routed through a steam-generating boiler (referred to as a CO boiler) where the carbon
monoxide in the flue gas is burned as fuel to provide steam for use in the refinery as well as to comply with any
applicable environmental regulatory limits on carbon monoxide emissions.[3]
The flue gas is finally processed through an electrostatic precipitator (ESP) to remove residual particulate matter to
comply with any applicable environmental regulations regarding particulate emissions. The ESP removes particulates
48 CHAPTER 6. FLUID CATALYTIC CRACKING

in the size range of 2 to 20 µm from the flue gas.[3] Particulate filter systems, known as Fourth Stage Separators (FSS)
are sometimes required to meet particulate emission limits. These can replace the ESP when particulate emissions
are the only concern.
The steam turbine in the flue gas processing system (shown in the above diagram) is used to drive the regenerator’s
combustion air compressor during start-ups of the FCC unit until there is sufficient combustion flue gas to take over
that task.

6.2 Chemistry

Before delving into the chemistry involved in catalytic cracking, it will be helpful to briefly discuss the composition
of petroleum crude oil.
Petroleum crude oil consists primarily of a mixture of hydrocarbons with small amounts of other organic compounds
containing sulfur, nitrogen and oxygen. The crude oil also contains small amounts of metals such as copper, iron,
nickel and vanadium.[2]
The elemental composition ranges of crude oil are summarized in Table 1 and the hydrocarbons in the crude oil can
be classified into three types:[1][2]

• Paraffins or alkanes: saturated straight-chain or branched hydrocarbons, without any ring structures

• Naphthenes or cycloalkanes: saturated hydrocarbons having one or more ring structures with one or more
side-chain paraffins

• Aromatics: hydrocarbons having one or more unsaturated ring structures such as benzene or unsaturated poly-
cyclic ring structures such as naphthalene or phenanthrene, any of which may also have one or more side-chain
paraffins.

Olefins or alkenes, which are unsaturated straight-chain or branched hydrocarbons, do not occur naturally in crude
oil.
In plain language, the fluid catalytic cracking process breaks large hydrocarbon molecules into smaller molecules
by contacting them with powdered catalyst at a high temperature and moderate pressure which first vaporizes the
hydrocarbons and then breaks them. The cracking reactions occur in the vapor phase and start immediately when the
feedstock is vaporized in the catalyst riser.
Figure 2 is a very simplified schematic diagram that exemplifies how the process breaks high boiling, straight-chain
alkane (paraffin) hydrocarbons into smaller straight-chain alkanes as well as branched-chain alkanes, branched alkenes
(olefins) and cycloalkanes (naphthenes). The breaking of the large hydrocarbon molecules into smaller molecules is
more technically referred to by organic chemists as scission of the carbon-to-carbon bonds.
As depicted in Figure 2, some of the smaller alkanes are then broken and converted into even smaller alkenes and
branched alkenes such as the gases ethylene, propylene, butylenes, and isobutylenes. Those olefinic gases are valuable
for use as petrochemical feedstocks. The propylene, butylene and isobutylene are also valuable feedstocks for certain
petroleum refining processes that convert them into high-octane gasoline blending components.
As also depicted in Figure 2, the cycloalkanes (naphthenes) formed by the initial breakup of the large molecules are
further converted to aromatics such as benzene, toluene, and xylenes, which boil in the gasoline boiling range and
have much higher octane ratings than alkanes.
In the cracking process carbon is also produced which gets deposited on the catalyst (catalyst coke). The carbon
formation tendency or amount of carbon in a crude or FCC feed is measured with methods such as Micro Carbon
Residue, Conradson Carbon Residue or Ramsbottom Carbon Residue.
By no means does Figure 2 include all the chemistry of the primary and secondary reactions taking place in the fluid
catalytic process. There are a great many other reactions involved. However, a full discussion of the highly technical
details of the various catalytic cracking reactions is beyond the scope of this article and can be found in the technical
literature.[1][2][3][4]
6.2. CHEMISTRY 49

Figure 2: Diagrammatic example of the catalytic cracking of petroleum hydrocarbons

6.2.1 Catalysts

Modern FCC catalysts are fine powders with a bulk density of 0.80 to 0.96 g/cm3 and having a particle size distribution
ranging from 10 to 150 µm and an average particle size of 60 to 100 μm.[9][10] The design and operation of an FCC
unit is largely dependent upon the chemical and physical properties of the catalyst. The desirable properties of an
FCC catalyst are:

• Good stability to high temperature and to steam

• High activity

• Large pore sizes

• Good resistance to attrition

• Low coke production


50 CHAPTER 6. FLUID CATALYTIC CRACKING

A modern FCC catalyst has four major components: crystalline zeolite, matrix, binder, and filler. Zeolite is the
primary active component and can range from about 15 to 50 weight percent of the catalyst. The zeolite used in
FCC catalysts is referred to as faujasite or as Type Y and is composed of silica and alumina tetrahedra with each
tetrahedron having either an aluminum or a silicon atom at the center and four oxygen atoms at the corners. It is a
molecular sieve with a distinctive lattice structure that allows only a certain size range of hydrocarbon molecules to
enter the lattice. In general, the zeolite does not allow molecules larger than 8 to 10 nm (i.e., 80 to 90 ångströms) to
enter the lattice.[9][10]
The catalytic sites in the zeolite are strong acids (equivalent to 90% sulfuric acid) and provide most of the catalytic
activity. The acidic sites are provided by the alumina tetrahedra. The aluminum atom at the center of each alumina
tetrahedra is at a +3 oxidation state surrounded by four oxygen atoms at the corners which are shared by the neigh-
boring tetrahedra. Thus, the net charge of the alumina tetrahedra is −1 which is balanced by a sodium ion during
the production of the catalyst. The sodium ion is later replaced by an ammonium ion, which is vaporized when the
catalyst is subsequently dried, resulting in the formation of Lewis and Brønsted acidic sites. In some FCC catalysts,
the Brønsted sites may be later replaced by rare earth metals such as cerium and lanthanum to provide alternative
activity and stability levels.[9][10]
The matrix component of an FCC catalyst contains amorphous alumina which also provides catalytic activity sites and
in larger pores that allows entry for larger molecules than does the zeolite. That enables the cracking of higher-boiling,
larger feedstock molecules than are cracked by the zeolite.
The binder and filler components provide the physical strength and integrity of the catalyst. The binder is usually
silica sol and the filler is usually a clay (kaolin).
Nickel, vanadium, iron, copper and other metal contaminants, present in FCC feedstocks in the parts per million
range, all have detrimental effects on the catalyst activity and performance. Nickel and vanadium are particularly
troublesome. There are a number of methods for mitigating the effects of the contaminant metals:[11][12]

• Avoid feedstocks with high metals content: This seriously hampers a refinery’s flexibility to process various
crude oils or purchased FCC feedstocks.
• Feedstock feed pretreatment: Hydrodesulfurization of the FCC feedstock removes some of the metals and also
reduces the sulfur content of the FCC products. However, this is quite a costly option.
• Increasing fresh catalyst addition: All FCC units withdraw some of the circulating equilibrium catalyst as spent
catalyst and replaces it with fresh catalyst in order to maintain a desired level of activity. Increasing the rate
of such exchange lowers the level of metals in the circulating equilibrium catalyst, but this is also quite a costly
option.
• Demetallization: The commercial proprietary Demet Process removes nickel and vanadium from the withdrawn
spent catalyst. The nickel and vanadium are converted to chlorides which are then washed out of the catalyst.
After drying, the demetallized catalyst is recycled into the circulating catalyst. Removals of about 95 percent
nickel removal and 67 to 85 percent vanadium have been reported. Despite that, the use of the Demet process
has not become widespread, perhaps because of the high capital expenditure required.
• Metals passivation: Certain materials can be used as additives which can be impregnated into the catalyst or
added to the FCC feedstock in the form of metal-organic compounds. Such materials react with the metal con-
taminants and passivate the contaminants by forming less harmful compounds that remain on the catalyst. For
example, antimony and bismuth are effective in passivating nickel and tin is effective in passivating vanadium.
A number of proprietary passivation processes are available and fairly widely used.

The major suppliers of FCC catalysts worldwide include Albemarle Corporation, W.R. Grace Company and BASF
Catalysts (formerly Engelhard). The price for lanthanum oxide used in fluid catalytic cracking has risen from $5 per
kilogram in early 2010 to $140 per kilogram in June 2011.[13]

6.3 History
The first commercial use of catalytic cracking occurred in 1915 when Almer M. McAfee of Gulf Refining Company
developed a batch process using aluminum chloride (a Friedel Crafts catalyst known since 1877) to catalytically crack
heavy petroleum oils. However, the prohibitive cost of the catalyst prevented the widespread use of McAfee’s process
at that time.[2][14]
6.3. HISTORY 51

In 1922, a French mechanical engineer named Eugene Jules Houdry and a French pharmacist named E.A. Prud-
homme set up a laboratory near Paris to develop a catalytic process for converting lignite coal to gasoline. Supported
by the French government, they built a small demonstration plant in 1929 that processed about 60 tons per day
of lignite coal. The results indicated that the process was not economically viable and it was subsequently shut
down.[15][16][17]
Houdry had found that Fuller’s earth, a clay mineral containing aluminosilicates, could convert oil derived from the
lignite to gasoline. He then began to study the catalysis of petroleum oils and had some success in converting vaporized
petroleum oil to gasoline. In 1930, the Vacuum Oil Company invited him to come to the United States and he moved
his laboratory to Paulsboro, New Jersey.
In 1931, the Vacuum Oil Company merged with Standard Oil of New York (Socony) to form the Socony-Vacuum
Oil Company. In 1933, a small Houdry unit processed 200 barrels per day (32 m3 /d) of petroleum oil. Because of
the economic depression of the early 1930s, Socony-Vacuum was no longer able to support Houdry’s work and gave
him permission to seek help elsewhere.
In 1933, Houdry and Socony-Vacuum joined with Sun Oil Company in developing the Houdry process. Three years
later, in 1936, Socony-Vacuum converted an older thermal cracking unit in their Paulsboro refinery in New Jersey
to a small demonstration unit using the Houdry process to catalytically crack 2,000 barrels per day (320 m3 /d) of
petroleum oil.
In 1937, Sun Oil began operation of a new Houdry unit processing 12,000 barrels per day (1,900 m3 /d) in their
Marcus Hook refinery in Pennsylvania. The Houdry process at that time used reactors with a fixed bed of catalyst and
was a semi-batch operation involving multiple reactors with some of the reactors in operation while other reactors
were in various stages of regenerating the catalyst. Motor-driven valves were used to switch the reactors between
online operation and offline regeneration and a cycle timer managed the switching. Almost 50 percent of the cracked
product was gasoline as compared with about 25 percent from the thermal cracking processes.[15][16][17]
By 1938, when the Houdry process was publicly announced, Socony-Vacuum had eight additional units under con-
struction. Licensing the process to other companies also began and by 1940 there were 14 Houdry units in operation
processing 140,000 barrels per day (22,000 m3 /d).
The next major step was to develop a continuous process rather than the semi-batch Houdry process. That step
was implemented by advent of the moving-bed process known as the Thermofor Catalytic Cracking (TCC) process
which used a bucket conveyor-elevator to move the catalyst from the regeneration kiln to the separate reactor section.
A small semicommercial demonstration TCC unit was built in Socony-Vacuum’s Paulsboro refinery in 1941 and
operated successfully, producing 500 barrels per day (79 m3 /d). Then a full-scale commercial TCC unit processing
10,000 barrels per day (1,600 m3 /d) began operation in 1943 at the Beaumont, Texas refinery of Magnolia Oil
Company, an affiliate of Socony-Vacuum. By the end of World War II in 1945, the processing capacity of the TCC
units in operation was about 300,000 barrels per day (48,000 m3 /d).
It is said that the Houdry and TCC units were a major factor in the winning of World War II by supplying the
high-octane gasoline needed by the air forces of Great Britain and the United States for the more efficient higher
compression ratio engines of the Spitfire and the Mustang.[15][16][17]
In the years immediately after World War II, the Houdriflow process and the air-lift TCC process were developed
as improved variations on the moving-bed theme. Just like Houdry’s fixed-bed reactors, the moving-bed designs
were prime examples of good engineering by developing a method of continuously moving the catalyst between the
reactor and regeneration sections. The first air-lift TCC unit began operation in October 1950 at the Beaumont, Texas
refinery.
This fluid catalytic cracking process had first been investigated in the 1920s by Standard Oil of New Jersey, but
research on it was abandoned during the economic depression years of 1929 to 1939. In 1938, when the success of
Houdry’s process had become apparent, Standard Oil of New Jersey resumed the project as part of a consortium of
that include five oil companies (Standard Oil of New Jersey, Standard Oil of Indiana, Anglo-Iranian Oil, Texas Oil
and Dutch Shell), two engineering-construction companies (M.W. Kellogg and Universal Oil Products) and a German
chemical company (I.G. Farben). The consortium was called Catalytic Research Associates (CRA) and its purpose
was to develop a catalytic cracking process which would not impinge on Houdry’s patents.[15][16][17]
Chemical engineering professors Warren K. Lewis and Edwin R. Gilliland of the Massachusetts Institute of Tech-
nology (MIT) suggested to the CRA researchers that a low velocity gas flow through a powder might “lift” it enough
to cause it to flow in a manner similar to a liquid. Focused on that idea of a fluidized catalyst, researchers Donald
Campbell, Homer Martin, Eger Murphree and Charles Tyson of the Standard Oil of New Jersey (now Exxon-Mobil
Company) developed the first fluidized catalytic cracking unit. Their U.S. Patent No. 2,451,804, A Method of and
52 CHAPTER 6. FLUID CATALYTIC CRACKING

Apparatus for Contacting Solids and Gases, describes their milestone invention. Based on their work, M. W. Kellogg
Company constructed a large pilot plant in the Baton Rouge, Louisiana refinery of the Standard Oil of New Jersey.
The pilot plant began operation in May 1940.
Based on the success of the pilot plant, the first commercial fluid catalytic cracking plant (known as the Model I FCC)
began processing 13,000 barrels per day (2,100 m3 /d) of petroleum oil in the Baton Rouge refinery on May 25, 1942,
just four years after the CRA consortium was formed and in the midst of World War II. A little more than a month
later, in July 1942, it was processing 17,000 barrels per day (2,700 m3 /d). In 1963, that first Model I FCC unit was
shut down after 21 years of operation and subsequently dismantled.[15][16][17][18]
In the many decades since the Model I FCC unit began operation, the fixed bed Houdry units have all been shut down
as have most of the moving bed units (such as the TCC units) while hundreds of FCC units have been built. During
those decades, many improved FCC designs have evolved and cracking catalysts have been greatly improved, but the
modern FCC units are essentially the same as that first Model I FCC unit.
Note: All of the refinery and company names in this history section (with the exception of Universal Oil Products)
have changed over time by mergers and buyouts. Some have changed a number of times.

6.4 See also


• Oil refinery

• Petroleum

• Cracking (chemistry)

• Catalysis

6.5 References
[1] James H. Gary and Glenn E. Handwerk (2001). Petroleum Refining: Technology and Economics (4th ed.). CRC Press.
ISBN 0-8247-0482-7.

[2] James. G. Speight (2006). The Chemistry and Technology of Petroleum (4th ed.). CRC Press. ISBN 0-8493-9067-2.

[3] Reza Sadeghbeigi (2000). Fluid Catalytic Cracking Handbook (2nd ed.). Gulf Publishing. ISBN 0-88415-289-8.

[4] David S.J. Jones and Peter P. Pujado (Editors) (2006). Handbook of Petroleum Processing (First ed.). Springer. ISBN
1-4020-2819-9.

[5] U.S. Downstream Processing of Fresh Feed Input by Catalytic Cracking Units (Energy Information Administration, U.S.
Dept. of Energy)

[6] Editorial Staff (November 2002). “Refining Processes 2002”. Hydrocarbon Processing: 108–112. ISSN 0887-0284.

[7] Fluid Catalytic Cracking

[8] Alex C. Hoffmann and Lewis E. Stein (2002). Gas Cyclones and Swirl Tubes:Principles, Design and Operation (1st ed.).
Springer. ISBN 3-540-43326-0.

[9] Jessica Elzea Kogel, Nikhil C. Trivedi, James M. Barber and Stanley T. Krukowsk (Editors) (2006). Industrial Minerals
& Rocks: Commodities, Markets and Uses (Seventh ed.). Society of Mining, Metallurgy and Exploration. ISBN 0-87335-
233-5.

[10] Wen-Ching Yang (2003). Handbook of Fluidization and Fluid Particle Systems. CRC Press. ISBN 0-8247-0259-X.

[11] Passivate Vanadium on FCC Catalysts for Improved Refinery Profitability (1997 Annual National Petrochemical and Re-
finers Association (NPRA) Meeting)

[12] Julius Scherzer (1990). Octane-enhancing Zeolitic FCC Catalysts: Scientific ans Technical Aspects. CRC Press. ISBN
0-8247-8399-9.

[13] Chu, Steven. Critical Materials Strategy page 17 United States Department of Energy, December 2011. Accessed: 23
December 2011.
6.6. EXTERNAL LINKS 53

[14] Pioneer of Catalytic Cracking: Almer McAfee at Gulf Oil (North American Catalysis Society website)

[15] Tim Palucka (Winter 2005). “The Wizard of Octane: Eugene Houdry”. Invention & Technology 20 (3).

[16] Amos A. Avidan, Michael Edwards and Hartley Owen (Mobil Research and Development) (January 8, 1990). “Innovative
Improvements Highlight FCC’s Past and Future”. Oil & Gas Journal 88 (2).

[17] “Houdry Process for Catalytic Cracking”. American Chemical Society. Retrieved April 27, 2012.

[18] Eger Murphree and the Four Horsemen: FCC, Fluid Catalytic Cracking (North American Catalysis Society website)

6.6 External links


• Valero Refinery Tour (Houston, TX) Description and diagram of power train

• CD Tech website discussion of Lummus FCC and hydrotreating of catalytically cracked naphtha.
• The FCC Network

• Recovery of CO from a FCC using the COPureSM Process

• North American Catalysis Society


• Fluid Catalytic Cracking (University of British Columbia, Quak Foo, Lee )

• CFD Simulation of a Full-Scale Commercial FCC Regenerator


Chapter 7

Cracking (chemistry)

In petroleum geology and chemistry, cracking is the process whereby complex organic molecules such as kerogens or
heavy hydrocarbons are broken down into simpler molecules such as light hydrocarbons, by the breaking of carbon-
carbon bonds in the precursors. The rate of cracking and the end products are strongly dependent on the temperature
and presence of catalysts. Cracking is the breakdown of a large alkane into smaller, more useful alkanes and alkenes.
Simply put, hydrocarbon cracking is the process of breaking a long-chain of hydrocarbons into short ones.
More loosely, outside the field of petroleum chemistry, the term “cracking” is used to describe any type of splitting
of molecules under the influence of heat, catalysts and solvents, such as in processes of destructive distillation or
pyrolysis.
Fluid catalytic cracking produces a high yield of gasoline and LPG, while hydrocracking is a major source of jet fuel,
Diesel fuel, naphtha, and LPG.

7.1 History and patents

Among several variants of thermal cracking methods (variously known as the "Shukhov cracking process", "Burton
cracking process", “Burton-Humphreys cracking process”, and “Dubbs cracking process”) Vladimir Shukhov, a Rus-
sian engineer, invented and patented the first in 1891 (Russian Empire, patent no. 12926, November 27, 1891).[1]
One installation was used to a limited extent in Russia, but development was not followed up. In the first decade of the
20th century the American engineers William Merriam Burton and Robert E. Humphreys independently developed
and patented a similar process as U.S. patent 1,049,667 on June 8, 1908. Among its advantages was the fact that
both the condenser and the boiler were continuously kept under pressure.[2]
In its earlier versions however, it was a batch process, rather than continuous, and many patents were to follow in the
USA and Europe, though not all were practical.[1] In 1924, a delegation from the American Sinclair Oil Corporation
visited Shukhov. Sinclair Oil apparently wished to suggest that the patent of Burton and Humphreys, in use by
Standard Oil, was derived from Shukhov’s patent for oil cracking, as described in the Russian patent. If that could be
established, it could strengthen the hand of rival American companies wishing to invalidate the Burton-Humphreys
patent. In the event Shukhov satisfied the Americans that in principle Burton’s method closely resembled his 1891
patents, though his own interest in the matter was primarily to establish that “the Russian oil industry could easily
build a cracking apparatus according to any of the described systems without being accused by the Americans of
borrowing for free”.[3]
At that time, just a few years after the Russian Revolution, Russia was desperate to develop industry and earn for-
eign exchange, so their oil industry eventually did obtain much of their technology from foreign companies, largely
American.[3] At about that time however, fluid catalytic cracking was being explored and developed and soon replaced
most of the purely thermal cracking processes in the fossil fuel processing industry. The replacement was however
not complete; many types of cracking, including pure thermal cracking, still are in use, depending on the nature of
the feedstock and the products required to satisfy market demands. Thermal cracking remains important however,
for example in producing naphtha, gas oil, and coke, and more sophisticated forms of thermal cracking have been
developed for various purposes. These include visbreaking, steam cracking, and coking.[4]

54
7.2. CHEMISTRY 55

Refinery using the Shukhov cracking process, Baku, Soviet Union, 1934.

7.2 Chemistry

A large number of chemical reactions take place during the cracking process, most of them based on free radicals.
Computer simulations aimed at modeling what takes place during steam cracking have included hundreds or even
thousands of reactions in their models. The main reactions that take place include:

7.2.1 Initiation

In these reactions a single molecule breaks apart into two free radicals. Only a small fraction of the feed molecules
actually undergo initiation, but these reactions are necessary to produce the free radicals that drive the rest of the
reactions. In steam cracking, initiation usually involves breaking a chemical bond between two carbon atoms, rather
than the bond between a carbon and a hydrogen atom.

CH3 CH3 → 2 CH3 •


56 CHAPTER 7. CRACKING (CHEMISTRY)

7.2.2 Hydrogen abstraction

In these reactions a free radical removes a hydrogen atom from another molecule, turning the second molecule into
a free radical.

CH3 • + CH3 CH3 → CH4 + CH3 CH2 •

7.2.3 Radical decomposition

In these reactions a free radical breaks apart into two molecules, one an alkene, the other a free radical. This is the
process that results in alkene products.

CH3 CH2 • → CH2 =CH2 + H•

7.2.4 Radical addition

In these reactions, the reverse of radical decomposition reactions, a radical reacts with an alkene to form a single,
larger free radical. These processes are involved in forming the aromatic products that result when heavier feedstocks
are used.

CH3 CH2 • + CH2 =CH2 → CH3 CH2 CH2 CH2 •

7.2.5 Termination

In these reactions two free radicals react with each other to produce products that are not free radicals. Two common
forms of termination are recombination, where the two radicals combine to form one larger molecule, and dispropor-
tionation, where one radical transfers a hydrogen atom to the other, giving an alkene and an alkane.

CH3 • + CH3 CH2 • → CH3 CH2 CH3


CH3 CH2 • + CH3 CH2 • → CH2 =CH2 + CH3 CH3

7.2.6 Example: cracking butane

There are three places where a butane molecule (CH3 -CH2 -CH2 -CH3 ) might be split. Each has a distinct likelihood:

• 48%: break at the CH3 -CH2 bond.

CH3 * / *CH2 -CH2 -CH3


Ultimately this produces an alkane and an alkene: CH4 + CH2 =CH-CH3

• 38%: break at a CH2 -CH2 bond.

CH3 -CH2 * / *CH2 -CH3


Ultimately this produces an alkane and an alkene of different types: CH3 -CH3 + CH2 =CH2

• 14%: break at a terminal C-H bond

H/CH2 -CH2 -CH2 -CH3


Ultimately this produces an alkene and hydrogen gas: CH2 =CH-CH2 -CH3 + H2
7.3. CRACKING METHODOLOGIES 57

7.3 Cracking methodologies

7.3.1 Thermal methods

Thermal cracking was the first category of hydrocarbon cracking to be developed. Thermal cracking is an example of
a reaction whose energetics are dominated by entropy (∆S°) rather than by enthalpy (∆H°) in the Gibbs Free Energy
equation ∆G°=∆H°-T∆S°. Although the bond dissociation energy D for a carbon-carbon single bond is relatively
high (about 375 kJ/mol) and cracking is highly endothermic, the large positive entropy change resulting from the
fragmentation of one large molecule into several smaller pieces, together with the extremely high temperature, makes
T∆S° term larger than the ∆H° term, thereby favoring the cracking reaction.

Thermal cracking

Modern high-pressure thermal cracking operates at absolute pressures of about 7,000 kPa. An overall process of
disproportionation can be observed, where “light”, hydrogen-rich products are formed at the expense of heavier
molecules which condense and are depleted of hydrogen. The actual reaction is known as homolytic fission and
produces alkenes, which are the basis for the economically important production of polymers.
Thermal cracking is currently used to “upgrade” very heavy fractions or to produce light fractions or distillates, burner
fuel and/or petroleum coke. Two extremes of the thermal cracking in terms of product range are represented by the
high-temperature process called “steam cracking” or pyrolysis (ca. 750 °C to 900 °C or higher) which produces
valuable ethylene and other feedstocks for the petrochemical industry, and the milder-temperature delayed coking
(ca. 500 °C) which can produce, under the right conditions, valuable needle coke, a highly crystalline petroleum coke
used in the production of electrodes for the steel and aluminium industries.
The first thermal cracking method, the Shukhov cracking process, was invented by Russian engineer Vladimir
Shukhov, in the Russian empire, Patent No. 12926, November 27, 1891.[5]
William Merriam Burton developed one of the earliest thermal cracking processes in 1912 which operated at 700–750
°F (371–399 °C) and an absolute pressure of 90 psi (620 kPa) and was known as the Burton process. Shortly thereafter,
in 1921, C.P. Dubbs, an employee of the Universal Oil Products Company, developed a somewhat more advanced
thermal cracking process which operated at 750–860 °F (399–460 °C) and was known as the Dubbs process.[6] The
Dubbs process was used extensively by many refineries until the early 1940s when catalytic cracking came into use.

Steam cracking

Steam cracking is a petrochemical process in which saturated hydrocarbons are broken down into smaller, often
unsaturated, hydrocarbons. It is the principal industrial method for producing the lighter alkenes (or commonly
olefins), including ethene (or ethylene) and propene (or propylene). Steam cracker units are facilities in which a
feedstock such as naphtha, liquefied petroleum gas (LPG), ethane, propane or butane is thermally cracked through
the use of steam in a bank of pyrolysis furnaces to produce lighter hydrocarbons. The products obtained depend on
the composition of the feed, the hydrocarbon-to-steam ratio, and on the cracking temperature and furnace residence
time.[7]
In steam cracking, a gaseous or liquid hydrocarbon feed like naphtha, LPG or ethane is diluted with steam and briefly
heated in a furnace without the presence of oxygen. Typically, the reaction temperature is very high, at around
850°C, but the reaction is only allowed to take place very briefly. In modern cracking furnaces, the residence time is
reduced to milliseconds to improve yield, resulting in gas velocities faster than the speed of sound. After the cracking
temperature has been reached, the gas is quickly quenched to stop the reaction in a transfer line heat exchanger or
inside a quenching header using quench oil.
The products produced in the reaction depend on the composition of the feed, the hydrocarbon to steam ratio and on
the cracking temperature and furnace residence time. Light hydrocarbon feeds such as ethane, LPGs or light naphtha
give product streams rich in the lighter alkenes, including ethylene, propylene, and butadiene. Heavier hydrocarbon
(full range and heavy naphthas as well as other refinery products) feeds give some of these, but also give products
rich in aromatic hydrocarbons and hydrocarbons suitable for inclusion in gasoline or fuel oil.
A higher cracking temperature (also referred to as severity) favors the production of ethene and benzene, whereas
lower severity produces higher amounts of propene, C4-hydrocarbons and liquid products. The process also results
in the slow deposition of coke, a form of carbon, on the reactor walls. This degrades the efficiency of the reactor, so
58 CHAPTER 7. CRACKING (CHEMISTRY)

reaction conditions are designed to minimize this. Nonetheless, a steam cracking furnace can usually only run for a
few months at a time between de-cokings. Decokes require the furnace to be isolated from the process and then a
flow of steam or a steam/air mixture is passed through the furnace coils. This converts the hard solid carbon layer to
carbon monoxide and carbon dioxide. Once this reaction is complete, the furnace can be returned to service.

7.3.2 Catalytic methods


The catalytic cracking process involves the presence of acid catalysts (usually solid acids such as silica-alumina and
zeolites) which promote a heterolytic (asymmetric) breakage of bonds yielding pairs of ions of opposite charges,
usually a carbocation and the very unstable hydride anion. Carbon-localized free radicals and cations are both highly
unstable and undergo processes of chain rearrangement, C-C scission in position beta as in cracking, and intra- and
intermolecular hydrogen transfer. In both types of processes, the corresponding reactive intermediates (radicals, ions)
are permanently regenerated, and thus they proceed by a self-propagating chain mechanism. The chain of reactions
is eventually terminated by radical or ion recombination.

Fluid Catalytic cracking

Main article: Fluid catalytic cracking


Fluid catalytic cracking is a commonly used process, and a modern oil refinery will typically include a cat cracker,

Schematic flow diagram of a fluid catalytic cracker

particularly at refineries in the US, due to the high demand for gasoline.[8][9][10] The process was first used around 1942
and employs a powdered catalyst. During WWII, in contrast to the Axis Forces which suffered severe shortages of
gasoline and artificial rubber, the Allied Forces were supplied with plentiful supplies of the materials. Initial process
7.4. SEE ALSO 59

implementations were based on low activity alumina catalyst and a reactor where the catalyst particles were suspended
in a rising flow of feed hydrocarbons in a fluidized bed.
Alumina-catalyzed cracking systems are still in use in high school and university laboratories in experiments concern-
ing alkanes and alkenes. The catalyst is usually obtained by crushing pumice stones, which contain mainly aluminium
oxide and silica into small, porous pieces. In the laboratory, aluminium oxide (or porous pot) must be heated.
In newer designs, cracking takes place using a very active zeolite-based catalyst in a short-contact time vertical or
upward-sloped pipe called the “riser”. Pre-heated feed is sprayed into the base of the riser via feed nozzles where
it contacts extremely hot fluidized catalyst at 1,230 to 1,400 °F (666 to 760 °C). The hot catalyst vaporizes the feed
and catalyzes the cracking reactions that break down the high-molecular weight oil into lighter components including
LPG, gasoline, and diesel. The catalyst-hydrocarbon mixture flows upward through the riser for a few seconds, and
then the mixture is separated via cyclones. The catalyst-free hydrocarbons are routed to a main fractionator for
separation into fuel gas, LPG, gasoline, naphtha, light cycle oils used in diesel and jet fuel, and heavy fuel oil.
During the trip up the riser, the cracking catalyst is “spent” by reactions which deposit coke on the catalyst and greatly
reduce activity and selectivity. The “spent” catalyst is disengaged from the cracked hydrocarbon vapors and sent to
a stripper where it is contacts steam to remove hydrocarbons remaining in the catalyst pores. The “spent” catalyst
then flows into a fluidized-bed regenerator where air (or in some cases air plus oxygen) is used to burn off the coke to
restore catalyst activity and also provide the necessary heat for the next reaction cycle, cracking being an endothermic
reaction. The “regenerated” catalyst then flows to the base of the riser, repeating the cycle.
The gasoline produced in the FCC unit has an elevated octane rating but is less chemically stable compared to other
gasoline components due to its olefinic profile. Olefins in gasoline are responsible for the formation of polymeric
deposits in storage tanks, fuel ducts and injectors. The FCC LPG is an important source of C3 -C4 olefins and
isobutane that are essential feeds for the alkylation process and the production of polymers such as polypropylene.

Hydrocracking

Hydrocracking is a catalytic cracking process assisted by the presence of added hydrogen gas. Unlike a hydrotreater,
where hydrogen is used to cleave C-S and C-N bonds, hydrocracking uses hydrogen to break C-C bonds (hydrotreat-
ment is conducted prior to hydrocracking to protect the catalysts in a hydrocracking).
The products of this process are saturated hydrocarbons; depending on the reaction conditions (temperature, pressure,
catalyst activity) these products range from ethane, LPG to heavier hydrocarbons consisting mostly of isoparaffins.
Hydrocracking is normally facilitated by a bifunctional catalyst that is capable of rearranging and breaking hydrocarbon
chains as well as adding hydrogen to aromatics and olefins to produce naphthenes and alkanes.
The major products from hydrocracking are jet fuel and diesel, but low sulphur naphta fractions and LPG are also
produced.[11] All these products have a very low content of sulfur and other contaminants.
It is very common in Europe and Asia because those regions have high demand for diesel and kerosene. In the US,
fluid catalytic cracking is more common because the demand for gasoline is higher.
The hydrocracking process depends on the nature of the feedstock and the relative rates of the two competing reac-
tions, hydrogenation and cracking. Heavy aromatic feedstock is converted into lighter products under a wide range
of very high pressures (1,000-2,000 psi) and fairly high temperatures (750°−1,500° F), in the presence of hydrogen
and special catalysts.
The primary function of hydrogen is, thus: a) If feedstock has a high paraffinic content, the primary function of
hydrogen is to prevent the formation of polycyclic aromatic compounds. b) Reduced tar formation c) Reduced
Impurities d) Prevent buildup of coke on the catalyst. e) High cetane fuel is achieved.

7.4 See also

• Fossil fuel reforming


60 CHAPTER 7. CRACKING (CHEMISTRY)

7.5 References
[1] M. S. Vassiliou (2 March 2009). Historical Dictionary of the Petroleum Industry. Scarecrow Press. pp. 459–. ISBN
978-0-8108-6288-3.

[2] Newton Copp; Andrew Zanella (1993). Discovery, Innovation, and Risk: Case Studies in Science and Technology. MIT
Press. pp. 172–. ISBN 978-0-262-53111-5.

[3] Oil of Russia. American Cracking for Soviet Refining. Yury Evdoshenko.

[4] Kraus, Richard S. Petroleum Refining Process in 78. Oil and Natural Gas, Kraus, Richard S., Editor, Encyclopedia of
Occupational Health and Safety, Jeanne Mager Stellman, Editor-in-Chief. International Labor Organization, Geneva. ©
2011.

[5] Vladimir Grigorievich Shukhov (Biography)

[6] U.S. Supreme Court Cases & Opinions, Volume 322, UNIVERSAL OIL PRODUCTS CO. V. GLOBE OIL & REFINING
CO., 322 U. S. 471 (1944)

[7] Propylene From Ethylene and Butene via Metathesis

[8] James H. Gary and Glenn E. Handwerk (2001). Petroleum Refining: Technology and Economics (4th ed.). CRC Press.
ISBN 0-8247-0482-7.

[9] James. G. Speight (2006). The Chemistry and Technology of Petroleum (4th ed.). CRC Press. ISBN 0-8493-9067-2.

[10] Reza Sadeghbeigi (2000). Fluid Catalytic Cracking Handbook (2nd ed.). Gulf Publishing. ISBN 0-88415-289-8.

[11] Sadighi, S., Ahmad, A., Shirvani, M. (2011) Comparison of lumping approaches to predict the product yield in a dual bed
VGO hydrocracker. , International Journal of Chemical Reactor Engineering, 9, art. no. A4.

7.6 External links


• Information on cracking in oil refining from howstuffworks.com
• Hydrocarbon Cracking — A Quick Summary for High School Students from canadaconnects.ca

• www.shukhov.org/shukhov.html — Vladimir Grigorievich Shukhov biography


Chapter 8

Visbreaker

A visbreaker is a processing unit in an oil refinery whose purpose is to reduce the quantity of residual oil produced
in the distillation of crude oil and to increase the yield of more valuable middle distillates (heating oil and diesel) by
the refinery. A visbreaker thermally cracks large hydrocarbon molecules in the oil by heating in a furnace to reduce
its viscosity and to produce small quantities of light hydrocarbons (LPG and gasoline).[1][2][3] The process name of
“visbreaker” refers to the fact that the process reduces (i.e., breaks) the viscosity of the residual oil. The process is
non-catalytic.

8.1 Process objectives

The objectives of visbreaking are:

• Reduce the viscosity of the feed stream: Typically this is the residue from vacuum distillation of crude oil but
can also be the residue from hydroskimming operations, natural bitumen from seeps in the ground or tar sands,
and even certain high viscosity crude oils.

• Reduce the amount of residual fuel oil produced by a refinery: Residual fuel oil is generally regarded as a low
value product. Demand for residual fuel continues to decrease as it is replaced in its traditional markets, such
as fuel needed to generate steam in power stations, by cleaner burning alternative fuels such as natural gas.

• Increase the proportion of middle distillates in the refinery output: Middle distillate is used as a diluent with
residual oils to bring their viscosity down to a marketable level. By reducing the viscosity of the residual stream
in a visbreaker, a fuel oil can be made using less diluent and the middle distillate saved can be diverted to higher
value diesel or heating oil manufacture.

8.2 Technology

8.2.1 Coil visbreaking

The term coil (or furnace) visbreaking is applied to units where the cracking process occurs in the furnace tubes
(or “coils”). Material exiting the furnace is quenched to halt the cracking reactions: frequently this is achieved by
heat exchange with the virgin material being fed to the furnace, which in turn is a good energy efficiency step, but
sometimes a stream of cold oil (usually gas oil) is used to the same effect. The gas oil is recovered and re-used. The
extent of the cracking reaction is controlled by regulation of the speed of flow of the oil through the furnace tubes.
The quenched oil then passes to a fractionator where the products of the cracking (gas, LPG, gasoline, gas oil and
tar) are separated and recovered.

61
62 CHAPTER 8. VISBREAKER

A schematic diagram of a Visbreaker unit

8.2.2 Soaker visbreaking


In soaker visbreaking, the bulk of the cracking reaction occurs not in the furnace but in a drum located after the
furnace called the soaker. Here the oil is held at an elevated temperature for a pre-determined period of time to
allow cracking to occur before being quenched. The oil then passes to a fractionator. In soaker visbreaking, lower
temperatures are used than in coil visbreaking. The comparatively long duration of the cracking reaction is used
instead.

8.2.3 Process options


Visbreaker tar can be further refined by feeding it to a vacuum fractionator. Here additional heavy gas oil may
be recovered and routed either to catalytic cracking, hydrocracking or thermal cracking units on the refinery. The
vacuum-flashed tar (sometimes referred to as pitch) is then routed to fuel oil blending. In a few refinery locations,
visbreaker tar is routed to a delayed coker for the production of certain specialist cokes such as anode coke or needle
coke.

8.2.4 Soaker visbreaking versus coil visbreaking


From the standpoint of yield, there is little or nothing to choose between the two approaches. However, each offers
significant advantages in particular situations:

• De-coking: The cracking reaction forms petroleum coke as a byproduct. In coil visbreaking, this deposits in
the tubes of the furnace and will eventually lead to fouling or blocking of the tubes. The same will occur in the
drum of a soaker visbreaker, though the lower temperatures used in the soaker drum lead to fouling at a much
slower rate. Coil visbreakers therefore require frequent de-coking. This is quite labour-intensive, but can be
developed into a routine where tubes are de-coked sequentially without the need to shutdown the visbreaking
8.3. QUALITY AND YIELDS 63

operation. Soaker drums require far less frequent attention but their being taken out of service normally requires
a complete halt to the operation. Which is the more disruptive activity will vary from refinery to refinery.

• Fuel Economy: The lower temperatures used in the soaker approach mean that these units use less fuel.
In cases where a refinery buys fuel to support process operations, any savings in fuel consumption could be
extremely valuable. In such cases, soaker visbreaking may be advantageous.

8.3 Quality and yields

8.3.1 Feed quality and product quality


The quality of the feed going into a visbreaker will vary considerably with the type of crude oil that the refinery is
processing. The following is a typical quality for the vacuum distillation residue of Arabian light (a crude oil from
Saudi Arabia and widely refined around the world):
Once this material has been run through a visbreaker (and, again, there will be considerable variation from visbreaker
to visbreaker as no two will operate under exactly the same conditions) the reduction in viscosity is dramatic:

8.3.2 Yields
The yields of the various hydrocarbon products will depend on the “severity” of the cracking operation as determined
by the temperature the oil is heated to in the visbreaker furnace. At the low end of the scale, a furnace heating to
425 °C would crack only mildly, while operations at 500 °C would be considered as very severe. Arabian light crude
residue when visbroken at 450 °C would yield around 76% (by weight) of tar, 15% middle distillates, 6% gasolines
and 3% gas and LPG.

8.3.3 Fuel oil stability


The severity of visbreaker operation is normally limited by the need to produce a visbreaker tar that can be blended
to make a stable fuel oil.
Stability in this case is taken to mean the tendency of a fuel oil to produce sediments when stored. These sediments
are undesirable as they can quickly foul the filters of pumps used to move the oil necessitating time-consuming
maintenance.
Vacuum residue fed to a visbreaker can be considered to be composed of the following:

• Asphaltenes: large polycyclic molecules that are suspended in the oil in a coloidal form

• Resins: also polycyclic but of a lower molecular weight than asphaltenes

• Aromatic hydrocarbons: derivatives of benzene, toluene and xylenes

• Parafinic hydrocarbons: alkanes

Visbreaking preferentially cracks aliphatic compounds which have relatively low sulphur contents, low density and
high viscosity and the effect of their removal can be clearly seen in the change in quality between feed and product. A
too severe cracking in a visbreaker will lead to the asphaltene colloid becoming metastable. Subsequent addition of
a diluent to manufacture a finished fuel oil can cause the colloid to break down, precipitating asphaltenes as a sludge.
It has been observed that a paraffinic diluent is more likely to cause precipitation than an aromatic one. Stability of
fuel oil is assessed using a number of proprietary tests (for example “P” value and SHF tests).

8.4 Economics
64 CHAPTER 8. VISBREAKER

8.4.1 Viscosity blending


The viscosity blending of two or more liquids having different viscosities is a three-step procedure. The first step is
to calculate the Viscosity Blending Index (VBI) of each component of the blend using the following equation (known
as a Refutas equation): [2][4]

(1) VBN = 14.534 × ln[ln(v + 0.8)] + 10.975

where v is the viscosity in square millimeters per second (mm²/s) or centistokes (cSt) and ln is the natural logarithm
(logₑ). It is important that the viscosity of each component of the blend be obtained at the same temperature.
The next step is to calculate the VBN of the blend, using this equation:

(2) VBNB ₑ = [wA × VBNA] + [wB × VBNB] + ... + [wX × VBNX]

where w is the weight fraction (i.e., % ÷ 100) of each component of the blend.
Once the viscosity blending number of a blend has been calculated using equation (2), the final step is to determine
the viscosity of the blend by using the invert of equation (1):

(VBN - 10.975) ÷ 14.534


(3) v = ee − 0.8

where VBN is the viscosity blending number of the blend and e is the transcendental number 2.71828, also known
as Euler’s number.

8.4.2 Example economics for a two-component blend


A marketable fuel oil, such as for fueling a power station, might be required to have a viscosity of 40 centistokes at
100 °C. It might be prepared using either the virgin or visbroken residue described above combined with a distillate
diluent (“cutter stock”). Such a cutter stock could typically have a viscosity at 100 °C of 1.3 centistokes. Rearranging
equation (2) above for a simple two component blend shows that the percentage of cutterstock required in the blend
is found by:
(4) %cutter stock = [VBN40 − VBNᵣₑ ᵢ ᵤₑ] ÷ [VBN ᵤ ₑᵣ ₒ − VBNᵣₑ ᵢ ᵤₑ]
Using the viscosities quoted in the tables above for the residues from Arab Light crude oil and calculating VBNs
according to equation (1) gives:
For virgin residue (i.e., the unconverted feed to the visbreaker): 27.5% cutter stock in the blend
For visbroken residue: 13.3% cutter stock in the blend.
As middle distillates have a far higher value in the market place than fuel oils, it can be seen that the use of a visbreaker
will considerably improve the economics of fuel oil manufacture. For example, if the cutter stock is taken to have
a value of $300 per tonne and fuel oil $150 per ton (oil prices naturally change quickly, but these prices, and more
importantly the differences between them, are not unrealistic), it is a simple matter to calculate the value of the
different residues in this example as being:
Virgin residue: $93.1 per tonne
Visbroken residue: $127.0 per tonne

8.5 References
[1] James H. Gary and Glenn E. Handwerk (1984). Petroleum Refining Technology and Economics (2nd Edition ed.). Marcel
Dekker, Inc. ISBN 0-8247-7150-8.

[2] Robert E. Maples (2000). Petroleum Refinery Process Economics (2nd Edition ed.). Pennwell Books. ISBN 0-87814-779-9.

[3] James G. Speight (2006). The Chemistry and Technology of Petroleum (4th Edition ed.). CRC Press. ISBN 0-8493-9067-2.

[4] C.T. Baird (1989), Guide to Petroleum Product Blending, HPI Consultants, Inc. HPI website
8.6. EXTERNAL LINKS 65

8.6 External links


• Foster Wheeler Visbreaking Overview

• Shell Thermal Conversion


• Shell Soaker Visbreaking

• Shell Deep Thermal Conversion

• Shell Thermal Gasoil Unit (TGU)


• Fuel Oil Stability Testing

• Bunkerworld: “Sediment Stability and Compatibility - The Structure of Fuel Oil”


Chapter 9

Merox

Merox is an acronym for mercaptan oxidation. It is a proprietary catalytic chemical process developed by UOP used
in oil refineries and natural gas processing plants to remove mercaptans from LPG, propane, butanes, light naphthas,
kerosene and jet fuel by converting them to liquid hydrocarbon disulfides.[1]
The Merox process requires an alkaline environment which, in some process versions, is provided by an aqueous
solution of sodium hydroxide (NaOH), a strong base, commonly referred to as caustic. In other versions of the
process, the alkalinity is provided by ammonia, which is a weak base.
The catalyst in some versions of the process is a water-soluble liquid. In other versions, the catalyst is impregnated
onto charcoal granules.
Processes within oil refineries or natural gas processing plants that remove mercaptans and/or hydrogen sulfide (H2 S)
are commonly referred to as sweetening processes because they results in products which no longer have the sour, foul
odors of mercaptans and hydrogen sulfide. The liquid hydrocarbon disulfides may remain in the sweetened products,
they may be used as part of the refinery or natural gas processing plant fuel, or they may be processed further.
Especially when dealing with kerosene, the Merox process is usually more economical than using a catalytic hydrodesulfurization
process for much the same purpose. Indeed, it is rarely (if ever) required to reduce the sulphur content of a straight-
run kerosene to respect the sulphur specification of Jet Fuel as the specification is 3000 ppm and very few crude oils
have a kerosene cut with a higher content of sulphur than this limit.

9.1 Types of Merox process units


UOP has developed many versions of the Merox process for various applications:

• Conventional Merox for extraction of mercaptans from LPG, propane, butanes or light naphthas.[2]

• Conventional Merox for sweetening jet fuels and kerosenes.[3]

• Merox for extraction of mercaptans from refinery and natural gases.[4]

• Minalk Merox for sweetening of naphthas.[5] This process continuously injects just a few ppm of caustic into
the feed naphtha.

• Caustic-free Merox for sweetening jet fuels and kerosenes.[6] This process injects small amounts of ammonia
and water (rather than caustic) into the feed naphtha to provide the required alkalinity.

• Caustic-free Merox for sweetening of naphthas.[7] This process also injects small amounts of ammonia and
water (rather than caustic) into the feed naphtha to provide the required alkalinity.

In all of the above Merox versions, the overall oxidation reaction that takes place in converting mercaptans to disulfides
is:

66
9.2. CONVENTIONAL MEROX FOR EXTRACTING MERCAPTANS FROM LPG 67

4 RSH + O2 → 2RSSR + 2H2 O

The most common mercaptans removed are:

• Methanethiol - CH3 SH [m-mercaptan]


• Ethanethiol - C2 H5 SH [e- mercaptan]
• 1-Propanethiol - C3 H7 SH [n-P mercaptan]
• 2-Propanethiol - CH3 CH(SH)CH3 [2C3 mercaptan]
• Butanethiol - C4 H9 SH [n-butyl mercaptan]
• tert-Butyl mercaptan - C(CH3 )3 SH [t-butyl mercaptan]
• Pentanethiol - C5 H11 SH [pentyl mercaptan]

In some of the above Merox process versions, the catalyst is a liquid. In others, the catalyst is in the form of impreg-
nated charcoal granules.
Process flow diagrams and descriptions of the two conventional versions of the Merox process are presented in the
following sections.

9.2 Conventional Merox for extracting mercaptans from LPG


The conventional Merox process for extraction and removal of mercaptans from liquefied petroleum gases (LPG),
such as propane, butanes and mixtures of propane and butanes, can also be used to extract and remove mercaptans
from light naphthas.[2] It is a two-step process. In the first step, the feedstock LPG or light naphtha is contacted in
the trayed extractor vessel with an aqueous caustic solution containing UOP’s proprietary liquid catalyst. The caustic
solution reacts with mercaptans and extracts them. The reaction that takes place in the extractor is:

2RSH + 2 NaOH → 2NaSR + 2 H2 O

In the above reaction, RSH is a mercaptan and R signifies an organic group such as a methyl, ethyl, propyl or other
group. For example, the ethyl mercaptan (ethanethiol) has the formula C2 H5 SH.
The second step is referred to as regeneration and it involves heating and oxidizing of the caustic solution leaving
the extractor. The oxidations results in converting the extracted mercaptans to organic disulfides (RSSR) which are
liquids that are water-insoluble and are then separated and decanted from the aqueous caustic solution. The reaction
that takes place in the regeneration step is:

4NaSR + O2 + 2H2 O → 2RSSR + 4NaOH

After decantation of the disulfides, the regenerated “lean” caustic solution is recirculated back to the top of the
extractor to continue extracting mercaptans.
The net overall Merox reaction covering the extraction and the regeneration step may be expressed as:

4 RSH + O2 → 2RSSR + 2H2 O

The feedstock entering the extractor must be free of any H2 S. Otherwise, any H2 S entering the extractor would
react with the circulating caustic solution and interfere with the Merox reactions. Therefore, the feedstock is first
“prewashed” by flowing through a batch of aqueous caustic to remove any H2 S. The reaction that takes place in the
prewash vessel is:

H2 S + NaOH → NaSH + H2 O

The batch of caustic solution in the prewash vessel is periodically discarded as "spent caustic" and replaced by fresh
caustic as needed.
68 CHAPTER 9. MEROX

9.2.1 Flow diagram


The flow diagram below depicts the equipment and the flow paths involved in the process.[2] The LPG (or light
naphtha) feedstock enters the prewash vessel and flows upward through a batch of caustic which removes any H2 S
that may be present in the feedstock. The coalescer at the top of the prewash vessel prevents caustic from being
entrained and carried out of the vessel.
The feedstock then enters the mercaptan extractor and flows upward through the contact trays where the LPG inti-
mately contacts the downflowing Merox caustic that extracts the mercaptans from the LPG. The sweetened LPG exits
the tower and flows through: a caustic settler vessel to remove any entrained caustic, a water wash vessel to further
remove any residual entrained caustic and a vessel containing a bed of rock salt to remove any entrained water. The
dry sweetened LPG exits the Merox unit.
The caustic solution leaving the bottom of the mercaptan extractor (“rich” Merox caustic) flows through a control valve
which maintains the extractor pressure needed to keep the LPG liquified. It is then injected with UOP’s proprietary
liquid catalyst (on an as needed basis), flows through a steam-heated heat exchanger and is injected with compressed
air before entering the oxidizer vessel where the extracted mercaptans are converted to disulfides. The oxidizer vessel
has a packed bed to keep the aqueous caustic and the water-insoluble disulfide well contacted and well mixed.
The caustic-disulfide mixture then flows into the separator vessel where it is allowed to form a lower layer of “lean”
Merox caustic and an upper layer of disulfides. The vertical section of the separator is for the disengagement and
venting of excess air and includes a Raschig ring section to prevent entrainment of any disulfides in the vented air.
The disulfides are withdrawn from the separator and routed to fuel storage or to a hydrotreater unit. The regenerated
lean Merox caustic is then pumped back to the top of the extractor for reuse.

9.3 Conventional Merox for sweetening jet fuel or kerosene


The conventional Merox process for the removal of mercaptans (i.e., sweetening) of jet fuel or kerosene is a one-
step process.[3] The mercaptan oxidation reaction takes place in an alkaline environment as the feedstock jet fuel or
kerosene, mixed with compressed air, flows through a fixed bed of catalyst in a reactor vessel. The catalyst consists
of charcoal granules that have been impregnated with UOP’s proprietary catalyst. The oxidation reaction that takes
place is:

4 RSH + O2 → 2RSSR + 2H2 O

As is the case with the conventional Merox process for treating LPG, the jet fuel or kerosene sweetening process also
requires that the feedstock be prewashed to remove any H2 S that would interfere with the sweetening. The reaction
that takes place in the batch caustic prewash vessel is:

H2 S + NaOH → NaSH + H2 O

9.3.1 Flow diagram


The Merox reactor is a vertical vessel containing a bed of charcoal granules that have been impregnated with the UOP
catalyst. The charcoal granules may be impregnated with the catalyst in situ or they may be purchased from UOP as
pre-impregnated with the catalyst. An alkaline environment is provided by caustic being pumped into reactor on an
intermittent, as needed basis.[3]
The jet fuel or kerosene feedstock from the top of the caustic prewash vessel is injected with compressed air and
enters the top of the Merox reactor vessel along with any injected caustic. The mercaptan oxidation reaction takes
place as the feedstock percolates downward over the catalyst. The reactor effluent flows through a caustic settler vessel
where it forms a bottom layer of aqueous caustic solution and an upper layer of water-insoluble sweetened product.
The caustic solution remains in the caustic settler so that the vessel contains a reservoir for the supply of caustic that
is intermittently pumped into the reactor to maintain the alkaline environment.
The sweetened product from the caustic settler vessel flows through a water wash vessel to remove any entrained
caustic as well as any other unwanted water-soluble substances, followed by flowing through a salt bed vessel to
remove any entrained water and finally through a clay filter vessel. The clay filter removes any oil-soluble substances,
9.4. SEE ALSO 69

Flow diagram of a conventional Merox process unit for extracting mercaptans from liquified petroleum gas (LPG).

organometallic compounds (especially copper) and particulate matter, which might prevent meeting jet fuel product
specifications.
The pressure maintained in the reactor is chosen so that the injected air will completely dissolve in the feedstock at
the operating temperature.

9.4 See also


• 2006 Côte d'Ivoire toxic waste spill

• Disulfide bond

9.5 References
[1] Treating Technology Solutions

[2] Merox Process for Mercaptan Extraction

[3] Merox Process for Kerosene/Jet Fuel Sweetening


70 CHAPTER 9. MEROX

Conventional Merox process unit for sweetening jet fuel or kerosene

[4] Merox Process for Gas Extraction

[5] Minalk Process for Fixed-Bed Naphtha Sweetening

[6] Caustic-free Merox Process for Kerosene/Jet Fuel Sweetening

[7] Caustic-Free Merox Process for Fixed-Bed Naphtha Sweetening

9.6 External links


• Energy and Environmental Profile of the U.S. Petroleum Refining Industry, U.S. Department of Energy Office
of Industrial Technologies, December 1998. (Use the keyword “Merox” in the PDF search function)
• Applications of the MeroxSM Process for Extraction of Mercaptan Sulfur Abstract of a paper presented at the
1999 Annual Meeting of the American Institute of Chemical Engineers.
• Profile of the Petroleum Refining Industry, Sector Notebook Project, Office of Enforcement and Compliance
Assurance, U.S. Environmental Protection Agency, September 1995. (Use the keyword “Merox” in the PDF
search function)
Chapter 10

Doctor sweetening process

Doctor Sweetening Process; version as patented by Kalinowsky (1954)

The Doctor Sweetening Process is an industrial chemical process for converting mercaptans in sour gasoline into
disulfides. Sulfur compounds darken gasoline, give it an offensive odor and increase toxic sulfur dioxide engine
emissions.[1] However, this process only reduces the odor.
These sulfur compounds can be removed with the following chemical reactions:[2]
(sour gasoline) 2RSH + Na2 PbO2 +S in the presence of NaOH ---- R-S-S-R + PbS + 2NaOH (alkyl disulfide)

10.1 Chemistry of the process


The chemistry of ‘doctor sweetening’ was described in detail by G. Wendt and S. Diggs in 1924. They showed that
the lead oxide solution brought about oxidation of the mercaptans to the corresponding organic disulfides, which

71
72 CHAPTER 10. DOCTOR SWEETENING PROCESS

are comparatively odourless. Lead oxide (litharge) will dissolve in reasonably concentrated solutions of sodium or
potassium hydroxide owing to formation of a soluble compound, sodium plumbite:

PbO + 2 NaOH −→ Na2 PbO2 + H2 O

When this alkaline solution is agitated with petroleum, the two liquids do not dissolve in one another, but any mer-
captan in the oil will unite with an equivalent amount of the lead (which then passes into the petroleum) to form what
is called a lead mercaptide, soluble in the oil:

2 RSH + Na2 PbO2 −→ (RS)2 Pb + 2 NaOH

If the mixture is now treated with powdered sulfur, which has a high affinity for lead, a black suspension of lead
sulfide forms, and conversion of the mercaptide into a so-called disulfide (which remains in the oil) is induced:

−(RS)2 Pb + S −→ RS − SR + PbS

With no sulfur added, but in the presence of atmospheric oxygen, the same conversion occurs, but only slowly, and
probably not completely:

2 (RS)2 Pb + 4 NaOH + O2 −→ 2 RS − SR + 2 Na2 PbO2 + 2 H2 O

It is evident that the process does not remove the sulfur from the oil but even may increase the sulphur content if too
much powdered sulfur is added, and some of the lead may remain in the petroleum.
The described chemistry is also the basis of the doctor test for the sweetness or sourness of gasoline (i.e., the extent
of sulfur contamination). A gasoline is described as doctor sweet if, after shaking with sodium plumbite solutions,
the addition of powdered sulfur fails to produce a dark precipitate of lead sulfide.

10.2 Literature
• McBryde, W.A.E.: Petroleum deodorized: Early canadian history of the ‘doctor sweetening’ process, Annals of
Science, Vol. 48, Issue 2, Taylor & Francis, 1991
• L. M. Henderson, W. B. Ross, C. M. Ridgway: Tetraethyllead Susceptibilities of Gasoline Doctor Treatment vs.
Caustic Washing, Ind. Eng. Chem., 1939, 31 (1), p. 27–30
• Naphtali, Max: Fortschritte auf dem Gebiete der Mineralöle. Die technische Entwicklung der Erdölindustrie
nach dem Kriege, Angewandte Chemie, Band 42, Ausgabe 20, p. 508-518, WILEY-VCH Verlag GmbH, May
18, 1929
• Otto Rotton, William Archer: Deodorizing Petroleum, American Artisan and Patent Record (New York), new
series 5, p. 310, 1867
• G.L. Wendt, S.H. Diggs: The Chemistry of “Sweetening” in the Petroleum Industry, Industrial and Engineering
Chemistry, Ausgabe 16, pp. 1113-1115, 1924
• M.L. Kalinowski: Doctor sweetening process using sulfur, US patent 2871187 January 27, 1957

10.3 References
[1] McGraw-Hill Dictionary of Scientific and Technical Terms, 6th edition, published by The McGraw-Hill Companies, Inc.,
2003

[2] G.L. Wendt, S.H. Diggs: The Chemistry of “Sweetening” in the Petroleum Industry, Industrial and Engineering Chemistry,
Ausgabe 16, p. 1113-1115, 1924
10.4. EXTERNAL LINKS 73

10.4 External links


• US patent 2871187, “Doctor sweetening process using sulfur”

• Saeid Mokhatab; William A. Poe; J. G. Speight (2006). Handbook of natural gas transmission and processing.
Gulf Professional Publishing. pp. 397–. ISBN 978-0-7506-7776-9. Retrieved 3 August 2010.
Chapter 11

Coker unit

A coker or coker unit is an oil refinery processing unit that converts the residual oil from the vacuum distillation
column or the atmospheric distillation column into low molecular weight hydrocarbon gases, naphtha, light and heavy
gas oils, and petroleum coke. The process thermally cracks the long chain hydrocarbon molecules in the residual oil
feed into shorter chain molecules leaving behind the excess carbon in the form of petroleum coke.
This petroleum coke can either be fuel grade (high in sulphur and metals) or anode grade (low in sulphur and metals).
The raw coke from the coker is often referred to as green coke.[1] In this context, “green” means unprocessed. The
further processing of green coke by calcining in a rotary kiln removes residual volatile hydrocarbons from the coke.
The calcined petroleum coke can be further processed in an anode baking oven in order to produce anode coke of
the desired shape and physical properties. The anodes are mainly used in the aluminium and steel industry.

11.1 Types of coker units


Main article: Delayed coker

There are three types of cokers used in oil refineries: Delayed coker, Fluid coker and Flexicoker.[2][3] The one that is
by far the most commonly used is the delayed coker.
The schematic flow diagram below depicts a typical delayed coker:

11.2 See also


• Delayed coker

• Shukhov cracking process

• Burton process

• Visbreaker

• Petroleum coke

11.3 References
[1] Petroleum coke on the website of the IUPAC Compendium of Chemical Terminology

[2] Gary, J.H. and Handwerk, G.E. (1984). Petroleum Refining Technology and Economics (2nd ed.). Marcel Dekker, Inc.
ISBN 0-8247-7150-8. OCLC 10323572.

[3] Hydrocarbon Refining staff (November 1998). “Refining Processes '98”. Hydrocarbon Processing: pages 62–64. ISSN
0887-0284.

74
11.4. EXTERNAL LINKS 75

A typical schematic flow diagram of a delayed coking unit

11.4 External links


• Detailed description of cokers and related topics
• Quality specifications for petroleum cokes
Chapter 12

Alkylation

Alkylation is the transfer of an alkyl group from one molecule to another. The alkyl group may be transferred as
an alkyl carbocation, a free radical, a carbanion or a carbene (or their equivalents).[1] Alkylating agents are widely
used in chemistry because the alkyl group is probably the most common group encountered in organic molecules.
Many biological target molecules or their synthetic precursors are composed of an alkyl chain with specific functional
groups in a specific order. Selective alkylation, or adding parts to the chain with the desired functional groups, is
used, especially if there is no commonly available biological precursor. Alkylation with only one carbon is termed
methylation.
In oil refining contexts, alkylation refers to a particular alkylation of isobutane with olefins. For upgrading of
petroleum, alkylation produces synthetic C7 –C8 alkylate, which is a premium blending stock for gasoline.[2]
In medicine, alkylation of DNA is used in chemotherapy to damage the DNA of cancer cells. Alkylation is accom-
plished with the class of drugs called alkylating antineoplastic agents.

Benzene Friedel-Crafts alkylation.

12.1 Alkylating agents

“Alkylating agent” redirects here. For the class of drugs, see alkylating antineoplastic agent.

Alkylating agents are classified according to their nucleophilic or electrophilic character.

12.1.1 Nucleophilic alkylating agents

Nucleophilic alkylating agents deliver the equivalent of an alkyl anion (carbanion). Examples include the use of
organometallic compounds such as Grignard (organomagnesium), organolithium, organocopper, and organosodium
reagents. These compounds typically can add to an electron-deficient carbon atom such as at a carbonyl group.
Nucleophilic alkylating agents can also displace halide substituents on a carbon atom. In the presence of catalysts,
they also alkylate alkyl and aryl halides, as exemplified by Suzuki couplings.

76
12.2. IN BIOLOGY 77

12.1.2 Electrophilic alkylating agents


Electrophilic alkylating agents deliver the equivalent of an alkyl cation. Examples include the use of alkyl halides
with a Lewis acid catalyst to alkylate aromatic substrates in Friedel-Crafts reactions. Alkyl halides can also react
directly with amines to form C-N bonds; the same holds true for other nucleophiles such as alcohols, carboxylic
acids, thiols, etc.
Electrophilic, soluble alkylating agents are often very toxic, due to their ability to alkylate DNA. They should be
handled with proper PPE. This mechanism of toxicity is also responsible for the ability of some alkylating agents to
perform as anti-cancer drugs in the form of alkylating antineoplastic agents, and also as chemical weapons such as
mustard gas. Alkylated DNA either does not coil or uncoil properly, or cannot be processed by information-decoding
enzymes. This results in cytotoxicity with the effects of inhibition the growth of the cell, initiation of programmed cell
death or apoptosis. However, mutations are also triggered, including carcinogenic mutations, explaining the higher
incidence of cancer after exposure.
Alcohols and phenols can be alkylated to give alkyl ethers:

R-OH + R'-X → R-O-R' + H-X

The produced acid HX is removed with a base, or, alternatively, the alcohol is deprotonated first to give an alkoxide
or phenoxide. For example, dimethyl sulfate alkylates the sodium salt of phenol to give anisole, the methyl ether of
phenol. The dimethyl sulfate is dealkylated to sodium methylsulfate.[3]

Ph-O– Na+ + Me2 SO4 → Ph-O-Me + Na+ MeSO4 –

On the contrary, the alkylation of amines introduces the problem that the alkylation of an amine makes it more
nucleophilic. Thus, when an electrophilic alkylating agent is introduced to a primary amine, it will preferentially
alkylate all the way to a quaternary ammonium cation.

R-NH2 → R-NH-R' → R-N(R')2 → R-N(R')3 + (alkylating agent omitted for clarity)

If the quaternary ammonium is not the desired product, more circuitious routes such as reductive amination are
necessary.

12.1.3 Carbene alkylating agents


Carbenes are extremely reactive and are known to attack even unactivated C-H bonds. Carbenes can be generated by
elimination of a diazo group. A metal can form a carbene equivalent called a transition metal carbene complex.
Trimethyloxonium tetrafluoroborate and triethyloxonium tetrafluoroborate belong to one of these categories.

12.2 In biology
Main article: methylation

Methylation is the most common type of alkylation, being associated with the transfer of a methyl group. Methylation
is distinct from alkylation in that it is specifically the transfer of one carbon, whereas alkylation can refer to the transfer
of long chain carbon groups. Methylation in nature is typically effected by vitamin B12-derived enzymes, where the
methyl group is carried by cobalt. In methanogenesis, coenzyme M is methylated by tetrahydromethanopterin.
Electrophilic compounds may alkylate different nucleophiles in the body. The toxicity, carcinogenity, and paradox-
ically, cancer cell-killing abilities of different DNA alkylating agents are an example.

12.3 Oil refining


Main article: alkylation unit
In a standard oil refinery process, isobutane is alkylated with low-molecular-weight alkenes (primarily a mixture
78 CHAPTER 12. ALKYLATION

Alkylation of alkenes (shown here is propene) by isobutane is a major process in refineries. It is catalysed by strong acids such as
HF and sulfuric acid.

of propene and butene) in the presence of a Bronsted acid catalyst, either sulfuric acid or hydrofluoric acid.[4] In
an oil refinery it is referred to as a sulfuric acid alkylation unit (SAAU) or a hydrofluoric alkylation unit, (HFAU).
Refinery workers may simply refer to it as the alky or alky unit. The catalyst protonates the alkenes (propene, butene)
to produce reactive carbocations, which alkylate isobutane. The reaction is carried out at mild temperatures (0 and
30 °C) in a two-phase reaction. Because the reaction is exothermic, cooling is needed: SAAU plants require lower
temperatures so the cooling medium needs to be chilled, for HFAU normal refinery cooling water will suffice. It is
important to keep a high ratio of isobutane to alkene at the point of reaction to prevent side reactions which produces a
lower octane product, so the plants have a high recycle of isobutane back to feed. The phases separate spontaneously,
so the acid phase is vigorously mixed with the hydrocarbon phase to create sufficient contact surface.
The product is called alkylate and is composed of a mixture of high-octane, branched-chain paraffinic hydrocarbons
(mostly isoheptane and isooctane). Alkylate is a premium gasoline blending stock because it has exceptional anti-
knock properties and is clean burning. Alkylate is also a key component of avgas. The octane number of the alkylate
depends mainly upon the kind of alkenes used and upon operating conditions. For example, isooctane results from
combining butylene with isobutane and has an octane rating of 100 by definition. There are other products in the
alkylate, so the octane rating will vary accordingly.
Since crude oil generally contains only 10 to 40 percent of hydrocarbon constituents in the gasoline range, refineries
use a fluid catalytic cracking process to convert high molecular weight hydrocarbons into smaller and more volatile
compounds, which are then converted into liquid gasoline-size hydrocarbons. Alkylation processes transform low
molecular-weight alkenes and iso-paraffin molecules into larger iso-paraffins with a high octane number.
Combining cracking, polymerization, and alkylation can result in a gasoline yield representing 70 percent of the
starting crude oil. More advanced processes, such as cyclicization of paraffins and dehydrogenation of naphthenes
forming aromatic hydrocarbons in a catalytic reformer, have also been developed to increase the octane rating of
gasoline. Modern refinery operation can be shifted to produce almost any fuel type with specified performance
criteria from a single crude feedstock.
Refineries examine whether it makes sense economically to install alkylation units. Alkylation units are complex,
with substantial economy of scale. In addition to a suitable quantity of feedstock, the price spread between the
value of alkylate product and alternate feedstock disposition value must be large enough to justify the installation.
Alternative outlets for refinery alklylation feedstocks include sales as LPG, blending of C4 streams directly into
gasoline and feedstocks for chemical plants. Local market conditions vary widely between plants. Variation in the
RVP specification for gasoline between countries and between seasons dramatically impacts the amount of butane
streams that can be blended directly into gasoline. The transportation of specific types of LPG streams can be
expensive so local disparities in economic conditions are often not fully mitigated by cross market movements of
alkylation feedstocks.
The availability of a suitable catalyst is also an important factor in deciding whether to build an alkylation plant. If
sulfuric acid is used, significant volumes are needed. Access to a suitable plant is required for the supply of fresh acid
and the disposition of spent acid. If a sulfuric acid plant must be constructed specifically to support an alkylation
unit, such construction will have a significant impact on both the initial requirements for capital and ongoing costs of
operation. Alternatively it is possible to install a WSA Process unit to regenerate the spent acid. No drying of the gas
takes place. This means that there will be no loss of acid, no acidic waste material and no heat is lost in process gas
reheating. The selective condensation in the WSA condenser ensures that the regenerated fresh acid will be 98% w/w
even with the humid process gas. It is possible to combine spent acid regeneration with disposal of hydrogen sulfide
by using the hydrogen sulfide as internal fuel in the refinery or elsewhere.[5]
The second main catalyst option is hydrofluoric acid. In typical alkylation plants, rates of consumption for acid are
much lower than for sulfuric acid. These plants also produce alkylate with better octane rating than do sulfuric plants.
However, due to its hazardous nature, HF acid is produced at very few locations and transportation must be managed
12.4. SEE ALSO 79

rigorously.

12.4 See also


• Hydrodealkylation
• Transalkylation

• Friedel–Crafts reaction

• Category:Alkylating agents
• Category:Ethylating agents
• Category:Methylating agents

12.5 References
[1] March Jerry; (1985). Advanced Organic Chemistry reactions, mechanisms and structure (3rd ed.). New York: John Wiley
& Sons, inc. ISBN 0-471-85472-7

[2] Stefanidakis, G.; Gwyn, J.E. (1993). “Alkylation”. In John J. McKetta. Chemical Processing Handbook. CRC Press. pp.
80–138. ISBN 0-8247-8701-3.

[3] G. S. Hiers and F. D. Hager (1941), “Anisole”, Org. Synth.; Coll. Vol. 1: 58

[4] Michael Röper, Eugen Gehrer, Thomas Narbeshuber, Wolfgang Siegel “Acylation and Alkylation” in Ullmann’s Encyclo-
pedia of Industrial Chemistry, Wiley-VCH, Weinheim, 2000. doi:10.1002/14356007.a01_185

[5] Sulphur recovery; (2007). The Process Principles, details advances in sulphur recovery by the WSA process. Denmark:
Jens Kristen Laursen, Haldor Topsøe A/S. Reprinted from Hydrocarbonengineering August 2007

12.6 External links


• Macrogalleria page on polycarbonate production

• Alkylating agents at the US National Library of Medicine Medical Subject Headings (MeSH)
Chapter 13

Isomerization

In chemistry isomerization (also isomerisation) is the process by which one molecule is transformed into another
molecule which has exactly the same atoms, but the atoms have a different arrangement e.g. A-B-C → B-A-C (these
related molecules are known as isomers [1] ). In some molecules and under some conditions, isomerization occurs
spontaneously. Many isomers are equal or roughly equal in bond energy, and so exist in roughly equal amounts,
provided that they can interconvert relatively freely, that is the energy barrier between the two isomers is not too high.
When the isomerization occurs intramolecularly it is considered a rearrangement reaction.
An example of an organometallic isomerization is the production of decaphenylferrocene, [(η5 -C5 Ph5 )2 Fe] from its
linkage isomer.[2][3]

13.1 Instances of isomerization


Isomerizations in hydrocarbon cracking is usually employed in organic chemistry, where fuels, such as pentane,
a straight-chain isomer, are heated in the presence of a platinum catalyst. The resulting mixture of straight- and
branched-chain isomers then have to be separated. An industrial process is also the isomerisation of n-butane into
isobutane.

“Trans-cis isomerism” is where in certain compounds an interconversion of cis and trans isomers can be observed, for
instance, with maleic acid and with azobenzene often by photoisomerization. Another example is the photochemical

80
13.2. SEE ALSO 81

conversion of the trans isomer to the cis isomer of resveratrol:[4]

Other instances are Aldose-ketose isomerism in biochemistry; isomerizations between conformational isomers, which
take place without an actual rearrangement for instance inconversion of two cyclohexane conformations; fluxional
molecules which display rapid interconversion of isomers e.g. bullvalene; and “valence isomerization": the isomer-
ization of molecules which involve structural changes resulting only from a relocation of single and double bonds. If
a dynamic equilibrium is established between the two isomers it is also referred to as valence tautomerism [5]
In a cycloisomerization a cyclic compound is formed.
The energy difference between two isomers is called “isomerization energy”. Isomerizations with low energy differ-
ence both experimental and computational (in parentheses) are endothermic trans-cis isomerization of 2-butene with
2.6 (1.2) kcal/mol, cracking of isopentane to n-pentane with 3.6 (4.0) kcal/mol or conversion of trans-2-butene to
1-butene with 2.6 (2.4) kcal/mol.[6]

13.2 See also


• Racemization

• Epimerization

13.3 References
[1] IUPAC, Compendium of Chemical Terminology, 2nd ed. (the “Gold Book”) (1997). Online corrected version: (2006–)
"isomerization".

[2] Brown, K. N.; Field, L. D.; Lay, P. A.; Lindall, C. M.; Masters, A. F. (1990). "(η5 -Pentaphenylcyclopentadienyl){1-
(η6 -phenyl)−2,3,4,5-tetraphenylcyclopentadienyl}iron(II), [Fe(η5 -C5 Ph5 ){(η6 -C6 H5 )C5 Ph4 }], a linkage isomer of de-
caphenylferrocene”. J. Chem. Soc., Chem. Commun. (5): 408–410. doi:10.1039/C39900000408.

[3] Field, L. D.; Hambley, T. W.; Humphrey, P. A.; Lindall, C. M.; Gainsford, G. J.; Masters, A. F.; Stpierre, T. G.; Webb, J.
(1995). “Decaphenylferrocene”. Aust. J. Chem. 48 (4): 851–860. doi:10.1071/CH9950851.

[4] Resveratrol Photoisomerization: An Integrative Guided-Inquiry Experiment Elyse Bernard, Philip Britz-McKibbin, Nicholas
Gernigon Vol. 84 No. 7 July 2007 Journal of Chemical Education 1159.

[5] Common Definitions and Terms in Organic Chemistry (from the website of Cartage.org.lb)

[6] How to Compute Isomerization Energies of Organic Molecules with Quantum Chemical Methods Stefan Grimme, Marc
Steinmetz, and Martin Korth J. Org. Chem.; 2007; 72(6) pp 2118 - 2126; (Article) doi:10.1021/jo062446p
Chapter 14

Amine gas treating

Amine gas treating, also known as amine scrubbing, gas sweetening and acid gas removal, refers to a group
of processes that use aqueous solutions of various alkylamines (commonly referred to simply as amines) to remove
hydrogen sulfide (H2 S) and carbon dioxide (CO2 ) from gases.[1][2][3] It is a common unit process used in refineries,
and is also used in petrochemical plants, natural gas processing plants and other industries.
Processes within oil refineries or chemical processing plants that remove hydrogen sulfide are referred to as “sweet-
ening” processes because the odor of the processed products is improved by the absence of hydrogen sulfide. An
alternative to the use of amines involves membrane technology. However, membranes are less attractive since the
relatively high capital and operation costs as well as other technical factors. [4]
Many different amines are used in gas treating:

• Diethanolamine (DEA)
• Monoethanolamine (MEA)
• Methyldiethanolamine (MDEA)
• Diisopropanolamine (DIPA)
• Aminoethoxyethanol (Diglycolamine) (DGA)

The most commonly used amines in industrial plants are the alkanolamines DEA, MEA, and MDEA. These amines
are also used in many oil refineries to remove sour gases from liquid hydrocarbons such as liquified petroleum gas
(LPG).

14.1 Description of a typical amine treater


Gases containing H2 S or both H2 S and CO2 are commonly referred to as sour gases or acid gases in the hydrocarbon
processing industries.
The chemistry involved in the amine treating of such gases varies somewhat with the particular amine being used.
For one of the more common amines, monoethanolamine (MEA) denoted as RNH2 , the chemistry may be expressed
as:

RNH2 + H2 S ⇔ RNH+
3 + SH-

A typical amine gas treating process (as shown in the flow diagram below) includes an absorber unit and a regenera-
tor unit as well as accessory equipment. In the absorber, the downflowing amine solution absorbs H2 S and CO2 from
the upflowing sour gas to produce a sweetened gas stream (i.e., a gas free of hydrogen sulfide and carbon dioxide)
as a product and an amine solution rich in the absorbed acid gases. The resultant “rich” amine is then routed into
the regenerator (a stripper with a reboiler) to produce regenerated or “lean” amine that is recycled for reuse in the
absorber. The stripped overhead gas from the regenerator is concentrated H2 S and CO2 .

82
14.2. AMINES 83

Process flow diagram of a typical amine treating process used in petroleum refineries, natural gas processing plants and other
industrial facilities.

14.2 Amines
The amine concentration in the absorbent aqueous solution is an important parameter in the design and operation of
an amine gas treating process. Depending on which one of the following four amines the unit was designed to use
and what gases it was designed to remove, these are some typical amine concentrations, expressed as weight percent
of pure amine in the aqueous solution:[1]

• Monoethanolamine: About 20 % for removing H2 S and CO2 , and about 32 % for removing only
CO2 .
• Diethanolamine: About 20 to 25 % for removing H2 S and CO2
• Methyldiethanolamine: About 30 to 55% % for removing H2 S and CO2
• Diglycolamine: About 50 % for removing H2 S and CO2

The choice of amine concentration in the circulating aqueous solution depends upon a number of factors and may be
quite arbitrary. It is usually made simply on the basis of experience. The factors involved include whether the amine
unit is treating raw natural gas or petroleum refinery by-product gases that contain relatively low concentrations of
both H2 S and CO2 or whether the unit is treating gases with a high percentage of CO2 such as the offgas from the
steam reforming process used in ammonia production or the flue gases from power plants.[1]
Both H2 S and CO2 are acid gases and hence corrosive to carbon steel. However, in an amine treating unit, CO2 is
the stronger acid of the two. H2 S forms a film of iron sulfide on the surface of the steel that acts to protect the steel.
84 CHAPTER 14. AMINE GAS TREATING

When treating gases with a high percentage of CO2 , corrosion inhibitors are often used and that permits the use of
higher concentrations of amine in the circulating solution.
Another factor involved in choosing an amine concentration is the relative solubility of H2 S and CO2 in the selected
amine.[1] The choice of the type of amine will affect the required circulation rate of amine solution, the energy
consumption for the regeneration and the ability to selectively remove either H2 S alone or CO2 alone if desired. For
more information about selecting the amine concentration, the reader is referred to Kohl and Nielsen’s book.
Activated MDEA or aMDEA uses piperazine as a catalyst to increase the speed of the reaction with CO2. It has been
commercially successful.[5]

14.3 Uses
In oil refineries, that stripped gas is mostly H2 S, much of which often comes from a sulfur-removing process called
hydrodesulfurization. This H2 S-rich stripped gas stream is then usually routed into a Claus process to convert it into
elemental sulfur. In fact, the vast majority of the 64,000,000 metric tons of sulfur produced worldwide in 2005 was
byproduct sulfur from refineries and other hydrocarbon processing plants.[6][7] Another sulfur-removing process is the
WSA Process which recovers sulfur in any form as concentrated sulfuric acid. In some plants, more than one amine
absorber unit may share a common regenerator unit. The current emphasis on removing CO2 from the flue gases
emitted by fossil fuel power plants has led to much interest in using amines for removing CO2 . (See also: Carbon
capture and storage and Conventional coal-fired power plant.)
In the specific case of the industrial synthesis of ammonia, for the steam reforming process of hydrocarbons to produce
gaseous hydrogen, amine treating is one of the commonly used processes for removing excess carbon dioxide in the
final purification of the gaseous hydrogen.

14.4 See also


• Ammonia production
• Hydrodesulfurization
• WSA Process
• Claus process
• Selexol
• Rectisol
• Amine
• Ionic liquids
• Solid sorbents for carbon capture

14.5 References
[1] Arthur Kohl and Richard Nielson (1997). Gas Purification (5th ed.). Gulf Publishing. ISBN 0-88415-220-0.

[2] Gary, J.H. and Handwerk, G.E. (1984). Petroleum Refining Technology and Economics (2nd ed.). Marcel Dekker, Inc.
ISBN 0-8247-7150-8.

[3] US 4080424, Loren N. Miller & Thomas S. Zawacki, “Process for acid gas removal from gaseous mixtures”, issued 21
Mar 1978, assigned to Institute of Gas Technology

[4] Baker, R. W. “Future Directions of Membrane Gas Separation Technology” Ind. Eng. Chem. Res. 2002, volume 41,
pages 1393-1411. doi:10.1021/ie0108088

[5] “Piperazine – Why It’s Used and How It Works”. The Contactor 2 (4). 2008. Retrieved 2013-10-23. |first1= missing
|last1= in Authors list (help)
14.6. EXTERNAL LINKS 85

[6] Sulfur production report by the United States Geological Survey

[7] Discussion of recovered byproduct sulfur

14.6 External links


• Description of Gas Sweetening Equipment and Operating Conditions
• Selecting Amines for Sweetening Units, Polasek, J. (Bryan Research & Engineering) and Bullin, J.A. (Texas
A&M University), Gas Processors Association Regional Meeting, Sept. 1994.
• Natural Gas Supply Association Scroll down to Sulfur and Carbon Dioxide Removal

• Description of the classic book on gas treating by Arthur Kohl & Richard Nielsen. Gas Purification (Fifth ed.).
Gulf Publishing. ISBN 0-88415-220-0.
Chapter 15

Claus process

Piles of sulfur produced in Alberta by the Claus process awaiting shipment at docks in Vancouver, Canada.

The Claus process is the most significant gas desulfurizing process, recovering elemental sulfur from gaseous
hydrogen sulfide. First patented in 1883 by the scientist Carl Friedrich Claus, the Claus process has become the
industry standard.
The multi-step Claus process recovers sulfur from the gaseous hydrogen sulfide found in raw natural gas and from the
by-product gases containing hydrogen sulfide derived from refining crude oil and other industrial processes. The by-
product gases mainly originate from physical and chemical gas treatment units (Selexol, Rectisol, Purisol and amine
scrubbers) in refineries, natural gas processing plants and gasification or synthesis gas plants. These by-product gases
may also contain hydrogen cyanide, hydrocarbons, sulfur dioxide or ammonia.
Gases with an H2 S content of over 25% are suitable for the recovery of sulfur in straight-through Claus plants while
alternate configurations such as a split-flow set up or feed and air preheating can be used to process leaner feeds.[1]
Hydrogen sulfide produced, for example, in the hydro-desulfurization of refinery naphthas and other petroleum oils,
is converted to sulfur in Claus plants.[2] The overall main reaction equation is:

2 H2 S + O2 → 2 S + 2 H2 O

In fact, the vast majority of the 64,000,000 metric tons of sulfur produced worldwide in 2005 was byproduct sul-
fur from refineries and other hydrocarbon processing plants.[3][4][5] Sulfur is used for manufacturing sulfuric acid,
medicine, cosmetics, fertilizers and rubber products. Elemental sulfur is used as fertilizer and pesticide.

86
15.1. HISTORY 87

15.1 History
The process was invented by Carl Friedrich Claus, a chemist working in England. A British patent was issued to him
in 1883. The process was later significantly modified by a German company called IG Farben[6]

15.2 Process description


A schematic process flow diagram of a basic 2+1-reactor (converter) SuperClaus unit is shown below:

Schematic flow diagram of a straight-through, 3 reactor (converter), Claus sulfur recovery unit.

The Claus technology can be divided into two process steps, thermal and catalytic.

15.2.1 Thermal step


In the thermal step, hydrogen sulfide-laden gas reacts in a substoichiometric combustion at temperatures above 850
°C [7] such that elemental sulfur precipitates in the downstream process gas cooler.
The H2 S content and the concentration of other combustible components (hydrocarbons or ammonia) determine the
location where the feed gas is burned. Claus gases (acid gas) with no further combustible contents apart from H2 S
are burned in lances surrounding a central muffle by the following chemical reaction:

2 H2 S + 3 O2 → 2 SO2 + 2 H2 O (ΔH = −4147.2 kJ mol−1 )

This is a strongly exothermic free-flame total oxidation of hydrogen sulfide generating sulfur dioxide that reacts away
in subsequent reactions. The most important one is the Claus reaction:

2 H2 S + SO2 → 3 S + 2 H2 O

The overall equation is:[5]

10 H2 S + 5 O2 → 2 H2 S + SO2 + 7/2 S2 + 8 H2 O

This equation shows that in the thermal step alone two-thirds of the hydrogen sulfide can be converted to sulfur.
Gases containing ammonia, such as the gas from the refinery’s sour water stripper (SWS), or hydrocarbons are con-
verted in the burner muffle. Sufficient air is injected into the muffle for the complete combustion of all hydrocarbons
and ammonia. The air to the acid gas ratio is controlled such that in total 1/3 of all hydrogen sulfide (H2 S) is converted
88 CHAPTER 15. CLAUS PROCESS

to SO2 . This ensures a stoichiometric reaction for the Claus reaction in the second catalytic step (see next section
below).
The separation of the combustion processes ensures an accurate dosage of the required air volume needed as a function
of the feed gas composition. To reduce the process gas volume or obtain higher combustion temperatures, the air
requirement can also be covered by injecting pure oxygen. Several technologies utilizing high-level and low-level
oxygen enrichment are available in industry, which requires the use of a special burner in the reaction furnace for this
process option.
Usually, 60 to 70% of the total amount of elemental sulfur produced in the process are obtained in the thermal process
step.
The main portion of the hot gas from the combustion chamber flows through the tube of the process gas cooler and is
cooled down such that the sulfur formed in the reaction step condenses. The heat given off by the process gas and the
condensation heat evolved are utilized to produce medium or low-pressure steam. The condensed sulfur is removed
at the liquid outlet section of the process gas cooler.
The sulfur forms in the thermal phase as highly reactive S2 diradicals which combine exclusively to the S8 allotrope:

4 S2 → S 8

Side reactions

Other chemical processes taking place in the thermal step of the Claus reaction are:[5]

• The formation of hydrogen gas:

2 H2 S → S2 + 2 H2 (ΔH > 0)
CH4 + 2 H2 O → CO2 + 4 H2

• The formation of carbonyl sulfide:

H2 S + CO2 → S=C=O + H2 O

• The formation of carbon disulfide:

CH4 + 2 S2 → S=C=S + 2 H2 S

15.2.2 Catalytic step


The Claus reaction continues in the catalytic step with activated aluminum(III) or titanium(IV) oxide, and serves to
boost the sulfur yield. More hydrogen sulfide (H2 S) reacts with the SO2 formed during combustion in the reaction
furnace in the Claus reaction, and results in gaseous, elemental sulfur.

2 H2 S + SO2 → 3 S + 2 H2 O (ΔH = −1165.6 kJ mol−1 )

This sulfur can be S6 , S7 , S8 or S9 .


The catalytic recovery of sulfur consists of three substeps: heating, catalytic reaction and cooling plus condensation.
These three steps are normally repeated a maximum of three times. Where an incineration or tail-gas treatment unit
(TGTU) is added downstream of the Claus plant, only two catalytic stages are usually installed.
The first process step in the catalytic stage is the gas heating process. It is necessary to prevent sulfur condensation in
the catalyst bed, which can lead to catalyst fouling. The required bed operating temperature in the individual catalytic
stages is achieved by heating the process gas in a reheater until the desired operating bed temperature is reached.
Several methods of reheating are used in industry:

• Hot-gas bypass: which involves mixing the two process gas streams from the process gas cooler (cold gas) and
the bypass (hot gas) from the first pass of the waste-heat boiler.
15.3. PROCESS PERFORMANCE 89

• Indirect steam reheaters: the gas can also be heated with high-pressure steam in a heat exchanger.
• Gas/gas exchangers: whereby the cooled gas from the process gas cooler is indirectly heated from the hot gas
coming out of an upstream catalytic reactor in a gas-to-gas exchanger.
• Direct-fired heaters: fired reheaters utilizing acid gas or fuel gas, which is burned substoichiometrically to avoid
oxygen breakthrough which can damage Claus catalyst.

The typically recommended operating temperature of the first catalyst stage is 315 °C to 330 °C (bottom bed temper-
ature). The high temperature in the first stage also helps to hydrolyze COS and CS2 , which is formed in the furnace
and would not otherwise be converted in the modified Claus process.
The catalytic conversion is maximized at lower temperatures, but care must be taken to ensure that each bed is
operated above the dew point of sulfur. The operating temperatures of the subsequent catalytic stages are typically
240 °C for the second stage and 200 °C for the third stage (bottom bed temperatures).
In the sulfur condenser, the process gas coming from the catalytic reactor is cooled to between 150 and 130 °C. The
condensation heat is used to generate steam at the shell side of the condenser.
Before storage, liquid sulfur streams from the process gas cooler, the sulfur condensers and from the final sulfur
separator are routed to the degassing unit, where the gases (primarily H2 S) dissolved in the sulfur are removed.
The tail gas from the Claus process still containing combustible components and sulfur compounds (H2 S, H2 and
CO) is either burned in an incineration unit or further desulfurized in a downstream tail gas treatment unit.

15.2.3 Sub dew point Claus process


The conventional Claus process described above is limited in its conversion due to the reaction equilibrium being
reached. Like all exothermic reactions, greater conversion can be achieved at lower temperatures, however as men-
tioned the Claus reactor must be operated above the sulfur dew point (120–150 °C) to avoid liquid sulfur physically
deactivating the catalyst. To overcome this problem, the sub dew point Claus process operates with reactors in paral-
lel. When one reactor has become saturated with adsorbed sulfur, the process flow is diverted to the standby reactor.
The reactor is then regenerated by sending process gas that has been heated to 300–350 °C to vaporize the sulfur.
This stream is sent to a condenser to recover the sulfur.

15.3 Process performance


Using two catalytic stages, the process will typically yield over 97% of the sulfur in the input stream. Over 2.6 tons
of steam will be generated for each ton of sulfur yield.
The physical properties of elemental sulfur obtained in the Claus process can differ from that obtained by other
processes.[5] Sulfur is usually transported as a liquid (melting point 115 °C). In ordinary sulfur viscosity can increase
rapidly at temperatures in excess of 160 °C due to the formation of polymeric sulfur chains but not so in Claus-sulfur.
Another anomaly is found in the solubility of residual H2 S in liquid sulfur as a function of temperature. Ordinarily
the solubility of a gas decreases with increasing temperature but now it is the opposite. This means that toxic and
explosive H2 S gas can build up in the headspace of any cooling liquid sulfur reservoir. The explanation for this
anomaly is the endothermic reaction of sulfur with H2 S to polysulfane.

15.4 See also


• Amine treating
• Hydro-desulfurization
• Crystasulf
• Hydrogenation
• Acid gas
• Sour gas
90 CHAPTER 15. CLAUS PROCESS

15.5 References
[1] Gas Processors Association Data Book, 10th Edition, Volume II, Section 22

[2] Gary, J.H. and Handwerk, G.E. (1984). Petroleum Refining Technology and Economics (2nd Edition ed.). Marcel Dekker,
Inc. ISBN 0-8247-7150-8.

[3] Sulfur production report by the United States Geological Survey

[4] Discussion of recovered byproduct sulfur

[5] Der Claus-Prozess. Reich an Jahren und bedeutender denn je, Bernhard Schreiner, Chemie in Unserer Zeit, Volume 42
Issue 6, Pages 378 - 392 2009

[6] Bibliographic Citation Sulfur Recovery Technology, B.G. Goar, American Institute of Chemical Engineers Spring National
Meeting, New Orleans, Louisiana, April 6, 1986

[7] Or between 950 and 1200 °C and even hotter near the flame, as stated in Der Claus-Prozess. Reich an Jahren und bedeu-
tender denn je, Bernhard Schreiner, Chemie in Unserer Zeit, Volume 42 Issue 6, 2009
15.6. TEXT AND IMAGE SOURCES, CONTRIBUTORS, AND LICENSES 91

15.6 Text and image sources, contributors, and licenses


15.6.1 Text
• Oil refinery Source: http://en.wikipedia.org/wiki/Oil%20refinery?oldid=639846896 Contributors: Bryan Derksen, Edward, Pit, Stan
Shebs, Theresa knott, Ender Everett, Bogdangiusca, Mulad, Gepwiki, Tpbradbury, Maximus Rex, SEWilco, Phil Boswell, Robbot,
Dale Arnett, RedWolf, Hemanshu, Pretzelpaws, BenFrantzDale, Tom harrison, Lupin, Leonard G., Unconcerned, Jason Quinn, Jorge
Stolfi, H Padleckas, AndrewKeenanRichardson, Scott Burley, Neutrality, Mike Rosoft, Yueni, Discospinster, Wk muriithi, Mani1, ES-
kog, *drew, Edward Z. Yang, BruceRD, Spalding, .:Ajvol:., Kjkolb, Nk, Hooperbloob, Jumbuck, Alansohn, Arthena, Andrewpmk, Ea-
gleamn, Snowolf, Dirac1933, Dan East, Psarianos, Xanthar, Nuno Tavares, PoccilScript, Firien, Rpalyvoda, Chris Buckey, Monk, Saper-
aud, Lhademmor, Coemgenus, Koavf, DRGrim, SeanMack, Ground Zero, Shimi925, Wavelength, Vuvar1, RussBot, Apancu, Ikar.us,
Stephenb, Salsb, Wimt, NawlinWiki, Wiki alf, AdiJapan, Stevenwmccrary58, 21655, Ageekgal, Wsiegmund, Stuhacking, Katieh5584,
robot, Luk, SmackBot, Saravask, KnowledgeOfSelf, Cdcon, Hardyplants, Knuto, Edgar181, Steam5, Gilliam, Hmains, Skizzik,
Bluebot, Sbharris, Colonies Chris, Antonrojo, Zsinj, Rogermw, Can't sleep, clown will eat me, Addshore, Kcordina, Markrobi, Falconeer,
Gbinal, DMacks, Nathanael Bar-Aur L., Anlace, John, CorvetteZ51, Aleator, Mbeychok, Boswell, Euphonisten, SimonD, Igoldste, Wood-
shed, Tawkerbot2, GBuilder, CmdrObot, Rorshacma, Dycedarg, Tex, Rdaveh, LouisBB, Rifleman 82, Lowellt, Nikopoley, M karzarj,
Hubba, MayaSimFan, Calvero JP, Aldis90, Richhoncho, Thijs!bot, Epbr123, Hazmat2, RobDe68, AntiVandalBot, Crentsch, Tillman,
Yellowdesk, David Shankbone, JAnDbot, Bongwarrior, VoABot II, Flavus Ahenobarbus, Dbushong, Kosm, Avicennasis, 28421u2232nfenfcenc,
Beagel, JaGa, JRWalko, Anonymous6494, Nono64, Preeteshdeora, Mausy5043, Tgeairn, J.delanoy, Bongomatic, Ali, AntSmith, Gurchzilla,
Hut 6.5, Lohrmann International, WilfriedC, Cometstyles, Useight, KGV, Idioma-bot, Funandtrvl, Spongessuck, VolkovBot, Lear’s Fool,
XterraGuy, Philip Trueman, LeaveSleaves, Raymondwinn, BotKung, Wtt, Ilyaroz, M0RD00R, Plazak, Meters, Samuel-em, Falcon8765,
Grsz11, Mugs2109, Allen Info, Caulde, Work permit, Erazlogo, Caltas, Lollypopcandygum, LeadSongDog, Happysailor, Oxymoron83,
Nuttycoconut, Lightmouse, Commutator, Alex.muller, Fratrep, Hatster301, OKBot, JL-Bot, Arinnian, ClueBot, Waynems, The Thing
That Should Not Be, Thehelpfulone, Thingg, Aitias, Putje81, DumZiBoT, Duckrob47, Lollypopcandygum1, XLinkBot, Thiswebsite-
suckssomuch, BodhisattvaBot, Dthomsen8, WikHead, Sicvolo, Thatguyflint, Elader, Addbot, Annwhent, Betterusername, Ronhjones,
Brandonlighty, MrOllie, Glane23, Lakshmix, Txag99, Tide rolls, Lightbot, Kahar13, Gail, JonathanBentz, Luckas-bot, 2D, Mauler90,
Fatal!ty, Daniele Pugliesi, Darolew, Flewis, Materialscientist, The High Fin Sperm Whale, Citation bot, Poofacejim, Sozchick101, Arthur-
Bot, BritishWatcher, J04n, Chubbz1995, Troyster87, Doulos Christos, Tech408, BNSDRX09, Insomnia64, Helloaday, FrescoBot, Pe-
terbmoorman, Aleksa Lukic, Pinethicket, Explocontrol, Utain, Allen Mesch, Aboriginal Noise, Hyd12, Sameeraemail, Thecoolmusic,
JimmyCraig, Vrenator, Minimac, Onel5969, RjwilmsiBot, TjBot, Ripchip Bot, Kiko4564, EmausBot, Francophile124, Desertroad,
Tommy2010, TeleComNasSprVen, Dh m83, Kittylionel, UrbanNerd, IGeMiNix, Donner60, Uthican, Scientific29, ChuispastonBot,
DASHBotAV, CharlieEchoTango, 28bot, Aaronw08, Zip822, ClueBot NG, Corusant, HISTMichelleS253, Widr, JohnSRoberts99, Cc-
cefalon, Arnavchaudhary, Mohamed CJ, M0rphzone, Northamerica1000, Mur61j, Silvio1973, Zujua, Grand Armor, Fozia nasir, Chris-
Gualtieri, YFdyh-bot, Asoleimanif, Geremy.Hebert, Makecat-bot, Frosty, Sfgiants1995, Arham pincha, Басилей, Robo4321, Adbar,
JaconaFrere, Ryanwilliams98, Amp121212, Michaelraymond40170, Marek Curda, Aquaman3000, Wcarias and Anonymous: 382
• Continuous distillation Source: http://en.wikipedia.org/wiki/Continuous%20distillation?oldid=599781823 Contributors: Giftlite, Bradeos
Graphon, H Padleckas, Alansohn, AzaToth, Walkerma, Vuo, Wsloand, Mindmatrix, Achim Raschka, Rjwilmsi, SeanMack, NielsenGW,
Edgar181, Colonies Chris, Mbeychok, Beetstra, RHB, LouisBB, Rifleman 82, Kozuch, Malleus Fatuorum, Thijs!bot, TimVickers, Ccr-
rccrr, Magioladitis, Rich257, WilfriedC, VolkovBot, TreasuryTag, Suprcel, Anonymous Dissident, Lightmouse, Fri117336, MystBot,
Addbot, MrOllie, Daniele Pugliesi, Citation bot, Fti74, Mnmngb, FrescoBot, Gire 3pich2005, Citation bot 1, Hyd12, ChemE50, Wagino
20100516, ClueBot NG, BarrelProof, MusikAnimal, Dave Bowman - Discovery Won, Tentinator and Anonymous: 16
• Vacuum distillation Source: http://en.wikipedia.org/wiki/Vacuum%20distillation?oldid=636180597 Contributors: Hpa, Mulad, Gentgeen,
Giftlite, Gtrmp, Revth, Rich Farmbrough, Kevin Dorner, Jope, Vuo, Wsloand, Drbreznjev, Mindmatrix, Tkslot, Rjwilmsi, Roboto de
Ajvol, Wiki alf, BOT-Superzerocool, SmackBot, Edgar181, Colonies Chris, Hgrosser, Cybercobra, AstroChemist, Mbeychok, Swot-
boy2000, BoH, Rifleman 82, Christian75, WinBot, Quantockgoblin, Jeepday, Emonkey, WilfriedC, VolkovBot, Dogsgomoo, Elsonhuge,
SieBot, Izmaelt, Kkolmetz, Pointillist, Addbot, JohnPaulGeorge, Bunnyhop11, Julia W, Daniele Pugliesi, Dinesh smita, Citation bot,
Elvim, ‫قلی زادگان‬, Mnmngb, Inhwiki, Pinethicket, Mys 721tx, Crimmm5, RjwilmsiBot, EmausBot, ZéroBot, ClueBot NG, VUGD,
Helpful Pixie Bot, Alkalinezeppelin, Pasicles, Jodosma, PREM KUMAR809809, Luca bonfanti and Anonymous: 41
• Hydrodesulfurization Source: http://en.wikipedia.org/wiki/Hydrodesulfurization?oldid=642026116 Contributors: Darkwind, Dragon-
flySixtyseven, R. S. Shaw, Polyparadigm, Rjwilmsi, Czar, Wavelength, RussBot, Hellbus, Light current, Bluebot, Cadmium, Smokefoot,
Heteren, Mion, Mbeychok, Beetstra, CmdrObot, Thijs!bot, OrenBochman, JAnDbot, Flavus Ahenobarbus, Beagel, User A1, DadaNeem,
Bob, Lamro, Elsonhuge, LeadSongDog, Wuhwuzdat, Shinkolobwe, XLinkBot, Addbot, Element16, Download, SpBot, Luckas-bot, Vy-
acheslav Nasretdinov, Daniele Pugliesi, Citation bot, Jü, Helloaday, Amse12, ZéroBot, Bromador, Shankk and Anonymous: 38
• Catalytic reforming Source: http://en.wikipedia.org/wiki/Catalytic%20reforming?oldid=636588869 Contributors: Robbot, Centrx, Michael
Devore, Unconcerned, Kjkolb, Arcenciel, Bhupesh mishra, Vuo, SeventyThree, Salsb, Tony1, Luk, SmackBot, Ma8thew, Knuto, Edgar181,
Mion, Mbeychok, JoeBot, Pro crast in a tor, Rich257, Beagel, STBot, R'n'B, WilfriedC, Bob, AlnoktaBOT, Mks004, Axiosaurus,
Lamro, Elsonhuge, ClueBot, Yikrazuul, XLinkBot, Addbot, Hornsofthebull, Eivindbot, Lightbot, Luckas-bot, Yobot, TaBOT-zerem,
AnomieBOT, Daniele Pugliesi, Citation bot, Xqbot, Hariehkr, Pderry, FoxBot, Nmillerche, EdoBot, Will Beback Auto, Silvio1973,
Hanan Halabi, Басилей and Anonymous: 29
• Fluid catalytic cracking Source: http://en.wikipedia.org/wiki/Fluid%20catalytic%20cracking?oldid=638925215 Contributors: Martin-
wguy, Bkonrad, Giraffedata, Kjkolb, Dziban303, Woohookitty, Wackyvorlon, Kolbasz, Bgwhite, Encyclops, Bug42, Ospalh, Open2universe,
Knuto, Amoore5000, Hmains, Chris the speller, TimBentley, Mbeychok, ChaoticLlama, OrenBochman, Faizhaider, Kosm, Loren-
zoB, Beagel, Hermann Luyken, WilfriedC, Lamro, Leebob69, Jojalozzo, Lightmouse, Rocksanddirt, Stillwaterising, BOTarate, Envi-
ronnement2100, XLinkBot, Dthomsen8, Salam32, Addbot, SPRich, Lightbot, OlEnglish, ‫ماني‬, Middayexpress, Luckas-bot, Daniele
Pugliesi, AdjustShift, Citation bot, Xqbot, Drilnoth, BeeJones, TobeBot, TGCP, EmausBot, John of Reading, ZéroBot, EdoBot, Ray-
ofway, ClueBot NG, KLindblom, Chemengr82, MerlIwBot, JohnSRoberts99, Klilidiplomus, Cool anurag upadhyay, CodedDeath, Kash-
may and Anonymous: 42
• Cracking (chemistry) Source: http://en.wikipedia.org/wiki/Cracking%20(chemistry)?oldid=640527849 Contributors: Magnus Manske,
Malcolm Farmer, Rmhermen, DrBob, Nonenmac, Liftarn, Egil, Mac, Glenn, Smack, Dysprosia, SEWilco, PuzzletChung, Robbot, COG-
DEN, Martinwguy, Giftlite, Unconcerned, Khalid hassani, Jlm255, Trif589, H Padleckas, Tsemii, Michall, GPoss, Kami, Spalding,
92 CHAPTER 15. CLAUS PROCESS

Robotje, Arcadian, Ejrrjs, Deryck Chan, Slipperyweasel, Ral315, Hooperbloob, Alansohn, Theodore Kloba, Velella, Gene Nygaard,
Unixer, EvilOverlordX, Emerson7, V8rik, Li-sung, Yurik, Rui Silva, Rangek, DVdm, Bgwhite, YurikBot, Hairy Dude, RussBot, Person
unknown, Hydrargyrum, Gaius Cornelius, Annabel, Stephen Burnett, Madcoverboy, THB, Tony1, Elkman, Wknight94, Orchid Righteous,
TBadger, DVD R W, Itub, Danielsavoiu, ThreeDee912, Slashme, Anastrophe, Hardyplants, Edgar181, Gilliam, Bluebot, MalafayaBot,
Bobomarch, Uthbrian, Can't sleep, clown will eat me, Bisected8, Rrburke, Kcordina, Smokefoot, DMacks, Turkystuffing, Mion, DA3N,
Jaganath, Mbeychok, Tuspm, Sergei Arsenyev, SweetNeo85, Courcelles, Tawkerbot2, Mr3641, Van helsing, ThinkOutsideTheTesseract,
Dgw, Rifleman 82, Tunheim, Calvero JP, Thijs!bot, Jaxsonjo, Bethan 182, Nonagonal Spider, Dtgriscom, FearedInLasVegas, Ufwuct,
AntiVandalBot, Ritabest, JAnDbot, MER-C, Anarchie76, Teh Minish, Staib, Freedomlinux, VoABot II, Jeff Dahl, JamesBWatson,
Charlesreid1, Bobkeyes, LorenzoB, Beagel, Calltech, BoulderDuck, MartinBot, R'n'B, Patrick.Fisher, Ananbo, Ryan Postlethwaite, Wil-
friedC, Bob, Rehaananana, Buttonwillow mckittrick, Bligit123, Latulla, Ashlandchemist, Maxipuchi, Dirkbb, Deep1989, Logan, SieBot,
OMCV, Alex9788, Flyer22, Radon210, Gordonofcartoon, Anchor Link Bot, Stillwaterising, ClueBot, The Thing That Should Not Be,
R000t, Boing! said Zebedee, Myself248, Kvan1st, PL290, Addbot, Dannyb1 5, Annwhent, Element16, Neodop, Lightbot, Pietrow,
Luckas-bot, Fraggle81, SwisterTwister, Nikonoff, AnomieBOT, Rubinbot, PhaseChanger, Citation bot, Xqbot, Elvim, ‫قلی زادگان‬,
Dougofborg, FrescoBot, Pepper, Riventree, Dyschunky, Citation bot 1, Pinethicket, Keayp01, Full-date unlinking bot, TBloemink,
Reaper Eternal, TheGrimReaper NS, Reach Out to the Truth, Hornlitz, EmausBot, John of Reading, Dcirovic, ZéroBot, SFigler, Ar-
ifbiosensor, JonRichfield, ClueBot NG, MelbourneStar, Harirameshvara, JohnSRoberts99, HMSSolent, NotWith, Zujua, Klilidiplomus,
Muskid, Mogism, AmericanLemming and Anonymous: 238
• Visbreaker Source: http://en.wikipedia.org/wiki/Visbreaker?oldid=626197378 Contributors: SebastianHelm, Fudoreaper, Unconcerned,
Leandros, Spalding, Gene Nygaard, Firien, Graham87, Rjwilmsi, Wavelength, RussBot, Dupz, Bluebot, Colonies Chris, Jaganath, Mbey-
chok, CmdrObot, Alaibot, 0x845FED, Beagel, GQsm, Malkinthecat, WilfriedC, Addbot, Daniele Pugliesi, Citation bot, LilHelpa, Esmu
Igors, ДмитрОст, Cccefalon and Anonymous: 15
• Merox Source: http://en.wikipedia.org/wiki/Merox?oldid=638478346 Contributors: Subsolar, Chowbok, Mindmatrix, V8rik, Rjwilmsi,
Edgar181, Aleem, Mbeychok, The-Pope, Robert Ullmann, Kosm, Beagel, Andy Dingley, Modgathwan, SieBot, Addbot, PlankBot,
AnomieBOT, Daniele Pugliesi, FrescoBot, Gamewizard71, PhnomPencil, Silvio1973, Macofe and Anonymous: 17
• Doctor sweetening process Source: http://en.wikipedia.org/wiki/Doctor%20sweetening%20process?oldid=570893584 Contributors: Selket,
Bearcat, Karn, V8rik, SirGrant, Malcolma, SmackBot, Arielco, Scientizzle, Beagel, Eminem.prashanth, Lamro, Canis Lupus, 1ForThe-
Money, Addbot, Saehrimnir, Helpful Pixie Bot, Cccefalon and Anonymous: 4
• Coker unit Source: http://en.wikipedia.org/wiki/Coker%20unit?oldid=606353274 Contributors: Michael Devore, H Padleckas, Pol098,
Crzrussian, Pietdesomere, SmackBot, Knuto, Mbeychok, MarshBot, Rich257, Kosm, Oceanflynn, Ilyaroz, AlleborgoBot, Versus22, Ad-
dbot, DOI bot, Daniele Pugliesi, Citation bot, ChrisGualtieri, Spyglasses, Monkbot and Anonymous: 6
• Alkylation Source: http://en.wikipedia.org/wiki/Alkylation?oldid=635523965 Contributors: SimonP, Mrwojo, Axel Driken, Edward,
Scurra, Fanghong, GPoss, Spalding, Pen1234567, Walkerma, Vuo, SeventyThree, Sherpajohn, V8rik, Rjwilmsi, YurikBot, RobotE,
RussBot, Gaius Cornelius, Yyy, David Berardan, Itub, SmackBot, Edgar181, UberMD, Jfreyre, Smokefoot, Mbeychok, JHunterJ, Fvas-
concellos, CmdrObot, A876, Robinatron, Rifleman 82, Christian75, Pro crast in a tor, JAnDbot, D99figge, Quietdaniel, DAGwyn, MiPe,
Beagel, Calltech, VolkovBot, Shaundakulbara, Ggenellina, Lamro, AlleborgoBot, Yikrazuul, Wandering fox, Addbot, RogerRoger1,
Adrian 1001, Luckas-bot, Yobot, Amirobot, AnomieBOT, Daniele Pugliesi, Citation bot, LilHelpa, Xqbot, Bharel, Elvim, Prari, Amse12,
Diannaa, RjwilmsiBot, EmausBot, Dcirovic, Thomasfgibsonmph, Silvio1973, Thepasta and Anonymous: 22
• Isomerization Source: http://en.wikipedia.org/wiki/Isomerization?oldid=635903432 Contributors: Tarquin, 168..., Astronautics, Vsmith,
Snowolf, Melaen, Gene Nygaard, V8rik, Bruce1ee, Gurch, Jrtayloriv, Physchim62, Jared Preston, JDnCoke, KnowledgeOfSelf, Smoke-
foot, SilkTork, Mbeychok, CmdrObot, Rifleman 82, Thijs!bot, Nonagonal Spider, Quantockgoblin, Pharaoh of the Wizards, LordAnu-
bisBOT, Snowbot, EdChem, Addbot, Yobot, Citation bot, Elvim, GrouchoBot, BenzolBot, DrilBot, Île flottante, Reallynca, RjwilmsiBot,
ClueBot NG, BG19bot, Wikitrolol, Epicgenius, Tentinator, Bigfoottoecheese, Tu178 and Anonymous: 29
• Amine gas treating Source: http://en.wikipedia.org/wiki/Amine%20gas%20treating?oldid=636569656 Contributors: H Padleckas, Rich
Farmbrough, Cacycle, Spalding, R. S. Shaw, Hooperbloob, Allen3, Fivemack, RussBot, SmackBot, Smokefoot, Goatchze, Mbeychok,
Magioladitis, JamesBWatson, Beagel, R'n'B, WilfriedC, Funkysapien, Outpostn, SieBot, ClueBot, Kuikentje, Jdavis41759, Thebearix,
Vianello, Addbot, Naudefjbot, Borvan53, AnomieBOT, Daniele Pugliesi, Citation bot, Riventree, Zrhodes, Amse12, RjwilmsiBot, New-
pointgas, GoingBatty, K6ka, Gertdam, ZéroBot, Crbn2020, Widr, Testem, Syzygy32, Blockyblock567 and Anonymous: 24
• Claus process Source: http://en.wikipedia.org/wiki/Claus%20process?oldid=621186474 Contributors: Charles Matthews, R. S. Shaw,
Pearle, Japanese Searobin, Boothy443, Woohookitty, Mindmatrix, Firien, V8rik, Bosquewiki, SmackBot, Chris the speller, Goatchze,
Mbeychok, Rifleman 82, Storm63640, Faigl.ladislav, MisterWeatherbee, Beagel, MartinBot, WilfriedC, STBotD, VolkovBot, Funkys-
apien, Otispa, Pjoef, SieBot, Arnobarnard, Correogsk, Shinkolobwe, UrsoBR, MystBot, Addbot, Yobot, Daniele Pugliesi, Citation bot,
Asuterisuku, Webmina, EmausBot, Joshramsey, Helpful Pixie Bot, BG19bot and Anonymous: 23

15.6.2 Images
• File:AlbertaSulfurAtVancouverBC.jpg Source: http://upload.wikimedia.org/wikipedia/commons/f/fa/AlbertaSulfurAtVancouverBC.
jpg License: CC SA 1.0 Contributors: English Wikipedia, original upload 10 July 2005 by Leonard G. Original artist: Leonard G.
• File:AmineTreating.png Source: http://upload.wikimedia.org/wikipedia/commons/f/f9/AmineTreating.png License: CC-BY-SA-3.0
Contributors: ? Original artist: ?
• File:Anacortes_Refinery_31911.JPG Source: http://upload.wikimedia.org/wikipedia/commons/5/51/Anacortes_Refinery_31911.JPG
License: CC BY 2.5 Contributors: Own work Original artist: Walter Siegmund (talk)
• File:Anacortes_Refinery_32017.JPG Source: http://upload.wikimedia.org/wikipedia/commons/2/2c/Anacortes_Refinery_32017.JPG
License: CC BY 2.5 Contributors: Own work Original artist: Walter Siegmund (talk)
• File:BP_PLANT_EXPLOSION-1_lowres2.jpg Source: http://upload.wikimedia.org/wikipedia/commons/7/71/BP_PLANT_EXPLOSION-1_
lowres2.jpg License: Public domain Contributors: http://www.csb.gov/assets/news/image/BP_PLANT_EXPLOSION-1_lowres2.jpg Orig-
inal artist: Chemical safety and hazards investigation board
• File:Benzene_Friedel-Crafts_alkylation-diagram.svg Source: http://upload.wikimedia.org/wikipedia/commons/f/f4/Benzene_Friedel-Crafts_
alkylation-diagram.svg License: CC BY 3.0 Contributors: Own work Original artist: Pen1234567
15.6. TEXT AND IMAGE SOURCES, CONTRIBUTORS, AND LICENSES 93

• File:CatReformer.png Source: http://upload.wikimedia.org/wikipedia/commons/2/21/CatReformer.png License: CC-BY-SA-3.0 Con-


tributors: I drew this flow diagram myself and currently I own all rights to it. I used Microsoft’s Paint program to draw it. I am
User:mbeychok and the date is December 5, 2006. Original artist: User:mbeychok
• File:CatReformerEq4.png Source: http://upload.wikimedia.org/wikipedia/commons/6/63/CatReformerEq4.png License: CC-BY-SA-
3.0 Contributors: ? Original artist: ?
• File:Catalytic_Reformer_Unit.jpg Source: http://upload.wikimedia.org/wikipedia/en/d/dd/Catalytic_Reformer_Unit.jpg License: ?
Contributors:
http://www.bp.com/popuppreviewthreecol.do?categoryId=121&contentId=7002769 Original artist:
British Petroleum p.l.c.
• File:Claus_Sulfur_Recovery.png Source: http://upload.wikimedia.org/wikipedia/commons/9/9f/Claus_Sulfur_Recovery.png License:
CC BY-SA 3.0 Contributors: Own work Original artist: Mbeychok
• File:Colonne_distillazione.jpg Source: http://upload.wikimedia.org/wikipedia/commons/c/cc/Colonne_distillazione.jpg License: CC
BY 3.0 Contributors: Own work Original artist: User:Luigi Chiesa
• File:Commons-logo.svg Source: http://upload.wikimedia.org/wikipedia/en/4/4a/Commons-logo.svg License: ? Contributors: ? Origi-
nal artist: ?
• File:Continuous_Binary_Fractional_Distillation.PNG Source: http://upload.wikimedia.org/wikipedia/commons/1/13/Continuous_
Binary_Fractional_Distillation.PNG License: CC-BY-SA-3.0 Contributors: ? Original artist: ?
• File:Continuous_Fractional_Distillation.PNG Source: http://upload.wikimedia.org/wikipedia/commons/0/08/Continuous_Fractional_
Distillation.PNG License: CC-BY-SA-3.0 Contributors: ? Original artist: ?
• File:ConvKeroMerox.png Source: http://upload.wikimedia.org/wikipedia/commons/1/12/ConvKeroMerox.png License: CC-BY-SA-
3.0 Contributors: ? Original artist: ?
• File:ConvLPGMerox.png Source: http://upload.wikimedia.org/wikipedia/commons/0/07/ConvLPGMerox.png License: CC-BY-SA-
3.0 Contributors: ? Original artist: ?
• File:Crude_Oil_Distillation.png Source: http://upload.wikimedia.org/wikipedia/commons/3/3c/Crude_Oil_Distillation.png License:
CC-BY-SA-3.0 Contributors: According to Theresa knott's user page (and her Image gallery), this image was originally created by her. -
Mbeychok 23:26, 20 September 2007 (UTC) Original artist: Users Psarianos, Theresa knott on en.wikipedia
• File:Crude_Oil_Distillation_Unit.png Source: http://upload.wikimedia.org/wikipedia/commons/9/94/Crude_Oil_Distillation_Unit.png
License: CC BY-SA 3.0 Contributors: Own work Original artist: Mbeychok
• File:Dehydrocyclization_reaction_of_heptane_to_toluene.svg Source: http://upload.wikimedia.org/wikipedia/commons/9/95/Dehydrocyclization_
reaction_of_heptane_to_toluene.svg License: Public domain Contributors: Own work Original artist: Yikrazuul (<a href='//commons.
wikimedia.org/wiki/User_talk:Yikrazuul' title='User talk:Yikrazuul'>talk</a>)
• File:Delayed_Coker.png Source: http://upload.wikimedia.org/wikipedia/commons/c/c1/Delayed_Coker.png License: Public domain
Contributors: Own work Original artist: Mbeychok
• File:DoctorSweeteningProcess-Kalinowsky.jpg Source: http://upload.wikimedia.org/wikipedia/commons/e/ea/DoctorSweeteningProcess-Kalinowsky.
jpg License: CC BY-SA 3.0 Contributors: Own work Original artist: Cccefalon
• File:ExxonMobil_Baton_Rouge.jpg Source: http://upload.wikimedia.org/wikipedia/commons/5/59/ExxonMobil_Baton_Rouge.jpg Li-
cense: CC BY-SA 3.0 Contributors: Own work Original artist: Adbar
• File:FCC.png Source: http://upload.wikimedia.org/wikipedia/commons/9/95/FCC.png License: Public domain Contributors: I drew this
myself using Microsoft’s paint program. Original artist: Mbeychok
• File:FCC_Chemistry.png Source: http://upload.wikimedia.org/wikipedia/commons/0/0b/FCC_Chemistry.png License: Public domain
Contributors: I drew this diagram myself using Microsoft’s Paint program Original artist: Mbeychok
• File:Fluid_Catalytic_Cracker.gif Source: http://upload.wikimedia.org/wikipedia/en/2/25/Fluid_Catalytic_Cracker.gif License: ? Con-
tributors:
http://www.secinfo.com/dsvrp.uEe6.d.htm#1stPage Original artist:
Valero Energy Corporation/TX
• File:Formation_of_decaphenylferrocene_from_its_linkage_isomer.PNG Source: http://upload.wikimedia.org/wikipedia/commons/
2/28/Formation_of_decaphenylferrocene_from_its_linkage_isomer.PNG License: Public domain Contributors: The uploader drew the
diagram using ChemOffice and the saved the image as a .PNG file using Microsoft Paint Original artist: user:EdChem
• File:HDS_Flow.png Source: http://upload.wikimedia.org/wikipedia/commons/6/6b/HDS_Flow.png License: GFDL Contributors: ?
Original artist: ?
• File:Haifa_Refinery_by_David_Shankbone.jpg Source: http://upload.wikimedia.org/wikipedia/commons/f/fa/Haifa_Refinery_by_
David_Shankbone.jpg License: CC-BY-SA-3.0 Contributors: David Shankbone Original artist: David Shankbone
• File:Hydrodesulfurization_cycle_for_thiophene_(simplified_diagram).png Source: http://upload.wikimedia.org/wikipedia/commons/
7/70/Hydrodesulfurization_cycle_for_thiophene_%28simplified_diagram%29.png License: Public domain Contributors: ? Original artist:
?
• File:IsobutaneAlkylation.png Source: http://upload.wikimedia.org/wikipedia/commons/7/7e/IsobutaneAlkylation.png License: CC BY-
SA 3.0 Contributors: Own work Original artist: Smokefoot
• File:Methylcyclohexanetotoluene.svg Source: http://upload.wikimedia.org/wikipedia/commons/7/7e/Methylcyclohexanetotoluene.svg
License: CC BY-SA 3.0 Contributors: Transferred from en.wikipedia; transferred to Commons by User:Ronhjones using CommonsHelper.
Original artist: Pderry (talk). Original uploader was Pderry at en.wikipedia
• File:N-pentane_isomerization.svg Source: http://upload.wikimedia.org/wikipedia/commons/1/15/N-pentane_isomerization.svg License:
Public domain Contributors: Transferred from en.wikipedia; transferred to Commons by User:Choij using CommonsHelper.
Original artist: V8rik (talk). Original uploader was V8rik at en.wikipedia
94 CHAPTER 15. CLAUS PROCESS

• File:Paraffintoisoparaffin.svg Source: http://upload.wikimedia.org/wikipedia/commons/f/f1/Paraffintoisoparaffin.svg License: CC


BY-SA 3.0 Contributors: Transferred from en.wikipedia; transferred to Commons by User:Ronhjones using CommonsHelper.
Original artist: Pderry (talk). Original uploader was Pderry at en.wikipedia
• File:Question_book-new.svg Source: http://upload.wikimedia.org/wikipedia/en/9/99/Question_book-new.svg License: Cc-by-sa-3.0
Contributors:
Created from scratch in Adobe Illustrator. Based on Image:Question book.png created by User:Equazcion Original artist:
Tkgd2007
• File:Rasveratrol_isomerization.png Source: http://upload.wikimedia.org/wikipedia/commons/7/74/Rasveratrol_isomerization.png Li-
cense: CC-BY-SA-3.0 Contributors: Transferred from en.wikipedia; transferred to Commons by User:Quadell using CommonsHelper.
Original artist: Original uploader was V8rik at en.wikipedia
• File:RefineryFlow.png Source: http://upload.wikimedia.org/wikipedia/commons/6/60/RefineryFlow.png License: CC-BY-SA-3.0 Con-
tributors: ? Original artist: ?
• File:Refinery_of_Slovnaft_in_Bratislava,_view_from_Nový_most_viewpoint.jpg Source: http://upload.wikimedia.org/wikipedia/
commons/0/07/Refinery_of_Slovnaft_in_Bratislava%2C_view_from_Nov%C3%BD_most_viewpoint.jpg License: CC BY 3.0 Contrib-
utors: Own work Original artist: Frettie
• File:Russian_Cracking.jpg Source: http://upload.wikimedia.org/wikipedia/commons/a/a5/Russian_Cracking.jpg License: Public do-
main Contributors: Trudy po istorii techniki, AN SSSR, 1954. Original artist: Unknown
• File:Simple_distillation_apparatus.svg Source: http://upload.wikimedia.org/wikipedia/commons/9/9c/Simple_distillation_apparatus.
svg License: Public domain Contributors: http://en.wikipedia.org/wiki/distillation Original artist: Original PNG by User:Quantockgoblin,
SVG adaptation by User:Slashme
• File:Symbol_support_vote.svg Source: http://upload.wikimedia.org/wikipedia/en/9/94/Symbol_support_vote.svg License: Public do-
main Contributors: ? Original artist: ?
• File:Tray_Distillation_Tower.PNG Source: http://upload.wikimedia.org/wikipedia/commons/e/e9/Tray_Distillation_Tower.PNG Li-
cense: CC BY-SA 2.5 Contributors: ? Original artist: ?
• File:Vacuum_Column.jpg Source: http://upload.wikimedia.org/wikipedia/commons/1/13/Vacuum_Column.jpg License: CC-BY-SA-
3.0 Contributors: The photograph was obtained from http://resources.schoolscience.co.uk/SPE/knowl/4/2index.htm?vacuum.html which
is a website of the Energy Institute of London, UK. Original artist: I do not know the name of the individual who actually took the
photograph.
• File:Vacuum_Distillation_Column.gif Source: http://upload.wikimedia.org/wikipedia/commons/2/28/Vacuum_Distillation_Column.
gif License: CC BY-SA 4.0 Contributors: Own work Original artist: This image has been created during “DensityDesign Integrated Course
Final Synthesis Studio” at Polytechnic University of Milan, organized by DensityDesign Research Lab. Image is released under CC-BY-SA
licence. Attribution goes to “Luca Bonfanti, DensityDesign Research Lab”.
• File:Vacuum_distillation_of_DMSO_at_70C.jpg Source: http://upload.wikimedia.org/wikipedia/commons/2/2c/Vacuum_distillation_
of_DMSO_at_70C.jpg License: Public domain Contributors: Transferred from en.wikipedia; Transfer was stated to be made by User:
Mbeychok. Original artist: Original uploader was Rifleman 82 at en.wikipedia
• File:Visbreaker.png Source: http://upload.wikimedia.org/wikipedia/en/5/58/Visbreaker.png License: Cc-by-sa-3.0 Contributors: ? Orig-
inal artist: ?
• File:Wiki_letter_w.svg Source: http://upload.wikimedia.org/wikipedia/en/6/6c/Wiki_letter_w.svg License: Cc-by-sa-3.0 Contributors:
? Original artist: ?

15.6.3 Content license


• Creative Commons Attribution-Share Alike 3.0

Potrebbero piacerti anche