Sei sulla pagina 1di 137

Restricted

Restricted

Deoiling
Manual

December2010
Restricted

Chapter 1. Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.1. Purpose of this manual . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.2. Document history . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.3. Organisation of the manual and changes from revision 1.1.. . . . . . . . . . . . . . . . . . . . . . . 12

2 Deoiling Manual V 2.0


Restricted

Chapter 2. Characterization of produced water. . . . . . . . . . . . . . . . . . . . . . . . . . . 13


2.1 General. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.2 When to characterize produced water. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.3 How to characterize produced water. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.3.1 Oil characterization for produced water treating. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.3.2 Biodegradation:. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.3.3 Hydrocarbons dispersed in water. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.3.4 Hydrocarbons dissolved in water. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.3.5 Dissolved organic carbon. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.3.6 Cations and anions (including heavy metals). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3.7 Salinity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3.8 pH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3.9 Hardness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3.10 Treatment chemicals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.3.11 Characterization of solids. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.3.12 Impact of solids on water treating. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.3.13 Solids of iron sulfides. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.3.14 Responding to solids problems. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.3.15 Examples of water treating problems caused by solids . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.3.16 Dissolved oxygen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.4 Droplet size distributions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.5 Characterization checklist and ranking. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.6 Examples of produced water characterization. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.6.1 North Sea versus Mars Deepwater Gulf of Mexico . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.6.2 Deepwater Gulf of Mexico example. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
Appendix 2.1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
Stokes’ law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
Appendix 2.1 Stokes’ law. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

Deoiling Manual V 2.0 3


Restricted

Chapter 3. Oil in water dispersions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39


3.1. General. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.2. Mixing intensity and maximum droplet size.. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.3. Mixing intensity and pressure drop. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
3.4. Droplet size distribution. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.5.Effect of pressure drop over a valve on droplet size distribution and separation. . . . . . . . . . 43
3.6. References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

4 Deoiling Manual V 2.0


Restricted

Chapter 4. Water disposal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45


4.1. General. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
4.2. Surface disposal – regulatory limits. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
4.3. Water injection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.3.1. General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
4.3.2. Water compatibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.3.3. Permeability impairment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.3.4. Solids removal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.3.5. Hydrocarbon removal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.3.6. Secondary waste water streams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
4.4. Disposal injection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
4.4.1. General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
4.4.2. Water quality constraints . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.5. SECONDARY (REJECT) STREAM DISPOSAL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.5.1. General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.5.2. Disposal options . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4.6. References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

Deoiling Manual V 2.0 5


Restricted

Chapter 5. Analysis and monitoring of waste water. . . . . . . . . . . . . . . . . . . . . . . . 53


5.1. General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
5.2. Oil in water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
5.3. Oil in water measurement – ir absorption methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.4. Oil in water measurement - other methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.4.1. Dispersed and dissolved hydrocarbons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.4.2. Gravimetric analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.4.3. Visible spectrum colorimetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
5.4.4. Gas chromatography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
5.4.5. Others . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
5.5. On-line oil-in-water analysers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.6. Measurement of droplet sizes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.6.1. General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.6.2. Sampling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.6.3. Potential problems in measuring droplet size distributions . . . . . . . . . . . . . . . . . . . . . . . . 58
5.6.4. Droplet size distribution analysers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
Appendix 5.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
Infrared absorption analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
A5.1.1 Sample preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
A5.1.2. Hydrocarbon extraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
A5.1.3 Infrared analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

6 Deoiling Manual V 2.0


Restricted

Chapter 6. Equipment selection and system integration . . . . . . . . . . . . . . . . . . . . . 67


6.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
6.2. Equipment selection and system design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
6.2.1. Sources and magnitude of water streams (1) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
6.2.2. Identify contaminants in the waste water stream (2) . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
6.2.3. Identify treated water quality requirements (3) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
6.2.4. Select a suitable process location for the water treatment system (4) . . . . . . . . . . . . . . . . 70
6.2.5. Identify upstream methods of improving ease of water treatment (5) . . . . . . . . . . . . . . . . . 70
6.2.6. Select the number of treatment stages (6) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
6.2.7. Select the suitable deoiling equipment (7) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
6.2.8. Treatment or disposal of secondary streams (8) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
6.2.9. System optimisation and integration (9) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
6.3. System optlmlsatlon and integration considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
6.3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
6.3.2. Maximising droplet size . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
6.3.4. Production separators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
6.3.5. Stable feed streams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
6.3.6. Recycle streams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
6.3.7. Mixing of water streams . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
6.3.8. Treatment chemicals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
6.4. Examples of water treatment schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
6.4.1. Draugen (Norway). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
6.4.2. Gamba (Gabon). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
6.4.3. NAM K2/G16 (Netherlands) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
6.4.4. Malampaya (Philippines). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
6.5 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79

Deoiling Manual V 2.0 7


Restricted

Chapter 7. Deoiling equipment – dispersed hydrocarbons . . . . . . . . . . . . . . . . . . . 81


7.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
7.2. Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
7.3. Coalescers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
7.4. Skimming tanks and vessels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
7.4.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
7.4.2. Performance variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
7.4.3. Configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
7.5 Discharge caissons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
7.5.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
7.5.2. Design guidelines Oil. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
7.5.3. Caisson internals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
7.5.4. Operational considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
7.6. API separator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
7.7. Plate interceptors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
7.7.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
7.7.2. Configurations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
7.7.3. Performance and sizing of plate packs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
7.7.4. Installation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
7.7.5. Operational considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
7.7.6. Concrete based standard CPI design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
7.8. Hydrocyclone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
7.8.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
7.8.2. Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
7.8.3. Performance variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
7.8.4. Installation/Configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
7.8.5. Operational considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
7.9. Induced gas flotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
7.9 .1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
7.9.2. Performance variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
7.9.3. Installation/Configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
7.9.5. Operational considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
7.10. Dissolved gas flotation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
7.10.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
7.10.2. Performance variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
7.10.3. Installation/configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
7.11. Deep bed media filtration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
7.11.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
7.11.2. Performance variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
7.11.3. Installation/Configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119
7.11.4. Design guidelines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
7.11.5. Operating experience . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
7.11.6. Nut shell deep bed media filters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
7.11.6.1. General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
7.11.6.2. Hydromation® . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
7.11.6.3. Wemco Silver Band . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
7.11.6.4 Walnut shell filter SIemens . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
7.11.6.5. Performance characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
7.11.6.6 Design guidelines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
7.11.6.7. Size and weight comparison . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
7.11.6.8. Application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
7.11.7. Absorption media filters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124
7.11.7.1 Performance characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124

8 Deoiling Manual V 2.0


Restricted

7.11.7.2. Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124


7.11.7.3. Design guidelines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
7.12. Membranes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
7.13. References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
Appendix 7.1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
API Separators. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
A7.1.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
A7.1.2. Performance variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
A7.1.3. Installation/Configuration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
A7.1.4. Design guidelines . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130

Deoiling Manual V 2.0 9


Restricted

Chapter 8. Deoiling equipment – dissolved hydrocarbons. . . . . . . . . . . . . . . . . . . 131


8.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
8.2. Levels of dissolved hydrocarbons in water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
8.3. Macro Porous Polymer Extraction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
8.3.1 General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132
8.3.2. Process description. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
8.3.3. Separation performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
8.3.4 Operating envelope . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135
8.3.5. Design guidelines. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
8.3.6. Process Location . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
8.3.7 Operational requirements. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
8.3.8. Veolia Services . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
8.3.9. Operating experience. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

10 Deoiling Manual V 2.0


Restricted

Chapter 1

Introduction

Deoiling Manual V 2.0 11


Restricted

1.1. Purpose of this manual


The main objectives of the Deoiling Manual are (1) to provide background information on all aspects of water
deoiling, including analysis and characterization, and the disposal of treated water, (2) to provide users with
an understanding of the design principles and performance characteristics of currently available equipment
for the removal of dispersed and dissolved hydrocarbons and (3) to provide guidelines for system design and
equipment selection.

1.2. Document history


The initial version of this manual was issued in January 1989 as part of the combined Dehydration/Deoiling
Manual. In 1993 it was withdrawn and the Deoiling Manual was revised and issued separately as Report
EP 93 – 1315. The current revision 2.0 is a further update by members of the EUHT Energy Utilities and
Heat Transfer group in Amsterdam, with the persistent support of Creative Services Amsterdam.

1.3. Organisation of the manual and changes from revision 1.1.


Chapters 2 (Characterization of Produced Water), 3 (Oil in water dispersions) have been rewritten com-
pletely.

Chapter 4 (Water disposal) has been actualized on the basis of the latest regulatory limits.
The chapter on Waste Water Sampling, previously chapter 5, has been superseded by the Sampling and
Analysis Guideline EP 2007-3186 and was taken out.

Chapter 5 (Analysis and Monitoring of waste water) was reorganized and partly rewritten in the light of the
changed legislation regarding regulatory tests and the phasing out of ozone depleting solvents. The in situ
microphotopgraphic droplet size analysis technique is no longer available, and has been taken out.

In chapter 6 (Equipment selection and system integration) new examples of water treating schemes were
included.

Chapter 7 (Deoiling equipment - dispersed hydrocarbons) has been limited to proven technology that is
actually deployed in E&P operations. The text has been made consistent with the DEP on Three-phase sepa-
rators (31.22.05.12) which has appeared after revision 1.1 was issued. Redundant descriptions of rotary
hydrocylones and centrifuges have been taken out, as well as detailed discussions of control schemes. New
information is provided about the optimization of skim tank internals by means of CFD simulations, the deoil-
ing capability of degassing vessels and more recent developments in media filtration and adsorption beds.

Chapter 8 (Deoiling equipment – dissolved hydrocarbons) has been limited to the currently preferred Macro
Porous Polymer Extraction. Gas stripping, activated carbon adsorption, solvent extraction and biotreatment
are not applied in E&P operations and have not been included.

12 Deoiling Manual V 2.0


Restricted

Chapter 2.

Characterization of produced
water

Deoiling Manual V 2.0 13


Restricted

2. Characterization of produced water


2.1 General
Each produced water stream is unique, with characteristics defined by a wide range of variables such as the
water source, processing operations and directly or indirectly added chemicals.

Correct characterization of a produced water stream is essential to ensure the appropriate selection, design
and operation of any deoiling system and eventual further processing. This chapter discusses a number of the
elements which define the characteristics of a waste water stream.

The primary goal is to provide practical advice on how to characterize produced water, when to character-
ize it, and what to do with the information that is obtained, in order to design a new facility or improve the
performance of an existing facility. For a checklist of what properties to include in a characterization, see
Table 1 below. Much of this chapter focuses on how to interpret the results and what to do with the informa-
tion in terms of equipment selection and design. These subjects are also discussed in later chapters.

The importance of proper characterization of produced water cannot be overstated. Nevertheless it is a


complicated task and most attempts to characterize produced water are deficient in one way or another.
While the technology available has improved dramatically over the past several years, there are still many
difficulties involved in the proper sampling, handling, analysis, and interpretation of results.

Each of the major steps involved in characterization have associated problems:


Sampling – poor sampling technique often changes the drop size distribution, and results in non-representa-
tive samples;
Transportation – the time between sampling and analysis often allows drops to coalesce;
Contamination – sample handling often allows oxygen to intrude or CO2 or H2S to escape;
Analysis – limitations of the analysis methods can lead to incorrect conclusions;
Interpretation – insufficient knowledge or experience which leads to erroneous conclusions;
Implementation – mistakes in detailed design which compromise overall design intent.
These problems are described further and practical advice about how to overcome them is given.

Sampling is not in any detail discussed because guidelines and manuals for sampling, sample handling and
how to avoid contamination already exist [1]. It is emphasized what should be analyzed, and what should
be done with the results, but not how to analyze. Again, guidelines and manuals for how to perform various
analyses are readily available [2 – 4].

In Section 2.3 a summary is given of the analyses that are used to characterize produced water. Since drop
size distribution is such an important part of produced water characterization, it is discussed in a separate
section (2.4). In Section 2.5 a checklist is given of analyses that should be carried out together with guidance
on how to use the results to improve system design or performance.

2.2 When to characterize produced water


The question of when to characterize produced water becomes a significant issue for a new project. On
deepwater floating facilities such as tension leg platforms and spars, separator vessel residence time is typi-
cally short due to the high cost of space and weight. On FPSOs deck space can be somewhat limited as well
but much less so than on TLP and Spars. Gas processing and liquid handling systems on such platforms are
typically designed with minimum spare capacity. During the design phase of such facilities, there is usually
little knowledge of the produced fluid properties. While the water treating system may not be required until
sometime after startup, space and weight must be allocated in the design phase.

Therefore, what is needed, and the guidance that is provided in this and further sections of this manual, is
practical advice on how to make judicious design decisions that build in flexibility at lowest cost and with
minimal space and weight requirements. In other words, the original system design must have sufficient built

14 Deoiling Manual V 2.0


Restricted

in flexibility to allow debottlenecking and capacity increase later when the produced water characteristics
are better known without major facility modifications. In fact, the timing of when to make design decisions is
one of the most important aspects of designing or debottlenecking a produced water system. That timing is
driven by the availability of information related to proper water characterization.

Even in the best case where water and oil samples have been taken at an early stage of the process design,
and have been carefully analyzed, such samples will not represent the drop size distribution of oil in water
which can only be assessed upon actual production. In such cases, water characterization should still be
carried out using the methods given here for characterizing the chemistry of the water and hydrocarbons. For
an estimate of drop size, theoretical considerations can be used, and analog fields and facilities should be
investigated for drop size.

2.3 How to characterize produced water


When considering produced water in E&P operations, the concentration of the produced hydrocarbons (oil in
water concentration) is normally the focus of both system design and environmental monitoring requirements.
However in addition to the dispersed hydrocarbons, produced water streams typically contain a wide variety
of contaminants, many of which have an impact on the effectiveness of the separation equipment, as well
as a potential impact on the environment. In order to minimize these effects, and in line with the Shell Group
commitment to identification and minimization of the waste streams, all the constituents of produced water
should be identified and accounted for.

The relative importance of contaminants in produced, discharge or disposal streams is dependent upon the
receiving process or environment. Environmental regulations vary depending on location. An example of
this is water salinity. Saline produced water may be relatively benign when discharged into a sea water
environment, but may have a significant environmental impact when discharged into a fresh water environ-
ment. Another example is in subsurface disposal (Produced Water Reinjection – PWRI) where produced
water characteristics must meet certain limits for oil and solids content in order to allow control of disposal
reservoir fracturing and fouling. A third example is where H2S scavenging is carried out using amine-based
chemistries. In that case, carbonate mineral scaling may occur unless prudent design decisions are made in
order to cater for the use of chemical injection and fluid monitoring.

To fully understand the characteristics of produced water, one must ultimately consider its source in the forma-
tion. In the formation, water is in equilibrium with hydrocarbon, and the formation rock material. As the water
is produced, the pressure and temperature decrease thus causing a shift in the equilibria. This may cause vari-
ous components to precipitate as solid particles. Other components may migrate to the oil/water interface. A
complete characterization of produced water will take into account the components of the co-produced gas
and oil phases, the composition of the reservoir rock, and phase equilibrium modeling will be carried out to
test for stability, consistency and to fill in any missing components.

Produced water can contain the following components:


■■ Important Components of Produced Water
■■ Dispersed oil
■■ Dissolved oil (hydrocarbons, BETX, phenols, PAH, etc)
■■ Dissolved organic acids (short chain fatty acids, naphthenic acids)
■■ Dissolved or precipitated minerals (NaCl, CaCO3, FeCO3, FeSx, BaSO4, SrSO4, etc)
■■ Dissolved metals (Fe, Zn, Cr, Mn, etc)
■■ Process & Production chemicals (corrosion inhibitor, water clarifier, methanol, glycols)
■■ Produced formation solids (clay, sand, carbonate)
■■ Dissolved and precipitated corrosion products (dissolved metals, solid metal oxides)
■■ Dissolved gases (O2, H2S, CO2)
■■ Combinations of the above (e.g. Schmoo)
■■ Various bacteria and by-products (SRB, GHB)

Deoiling Manual V 2.0 15


Restricted

The above list is given from the perspective of chemical entities, rather than types of analyses. Many of these
entities are determined in typical or standard analyses. The table below gives the recommended checklist for
water characteristics required for design and troubleshooting.

Table 1. Summary Checklist of Water Characteristics required for design and troubleshooting.

Oil droplet size distribution


Oil in water concentration
Oil flow assurance characterization (SARA, stability of wax, paraffin, asphaltenes)
Gas composition and oil characterization (H2S, CO2, C1 – C7, BTEX, TAN, biomarkers)
Water analysis (anions, cations, pH, hardness, alkalinity, dissolved gases & organics, COD, etc)
Suspended solids concentration, particle size distribution, composition and mineralogy
Desktop settling, visual observations, and optical microscopy

The material below discusses the typical ranges of measured values, and indications of problems, and how
to interpret the results of such a characterization.

2.3.1 Oil characterization for produced water treating


A typical produced water stream from an E&P operation will contain a wide variety of organic compounds
derived from contact with the produced oil. These organic compounds can be broadly grouped as dispersed
oil, dissolved hydrocarbons, dissolved polar organics, and dissolved gases. The relative importance of the
components depends on the requirements regarding discharge and disposal, as well as the environmental
impact, the potential for sustained biological activity in the case of waterflooding, and the effluent quality
required for reuse options. Hydrocarbons in water are considered to be dissolved if they pass through a 0.45
micron filter. Solids or dispersed oil in water is that which is retained on the filter.

Oil characterization is important for proper characterization of produced water because there are many
components in the hydrocarbon phase that contribute to the stability of oil in water drops. Organic acids,
asphaltenes, resins, are just some of the components that are found in the oil phase which can transfer to
water phase or to the oil/water interface and thus have an impact on produced water quality.

The simplest oil characterization parameter is the density or API gravity. Light gravity oils have a greater
density difference with water which thus helps drive gravity settling. They also have lower surface elasticity
which promotes drop coalescence. But their lower viscosity promotes drop break up in a shear zone. One
of the most useful characterizations of produced fluids is to calculate the Stokes factor on the basis of oil
density, water density, and water viscosity. The Stokes Factor S, is defined as S = Dr/m, where Dr is the
density difference between water and oil, and m the water viscosity. S lumps the physical properties which
determine the rising velocity of an oil droplet dispersed in water. The velocity can be established from Stokes
law when the droplet size is known. Stokes law is dealt with in more detail in Appendix 2.1. The Stokes
Factor is a measure of the settling tendency of oil in water, without taking into account the drop size. The
higher the Stokes Factor, the greater the speed with which oil drops rise to the surface of the water phase in
the separator vessel. It is one of the characterizations that can be carried out in the design phase before drop
size is known. An example of this is given in the Examples section.

The modified IP 143 / 57 method is Shell’s proprietary procedure to determine the SARA (saturates, aromat-
ics, resins, asphaltene) composition of a hydrocarbon sample. The method consists of an initial topping step,
followed by asphaltene precipitation and gravimetric analysis, and liquid chromatography to determine the
concentration of the remaining components. In order to determine the stability of asphaltenes, and thus deter-
mine their tendency to help stabilize oil in water emulsions, the results of a SARA analysis must be compared
with analog data and field experience. Generally, the ratio of asphaltene to resin and aromatics to saturates
gives an indication of stability. Further details are available elsewhere [5].

16 Deoiling Manual V 2.0


Restricted

2.3.2 Biodegradation:
Oil analysis should include Total Acid Number (TAN) which provides a simple first indication of the degree
of biodegradation. The recommended procedure is ASTM D-664. Besides TAN, organic acid content and
composition should also be measured. The definitive test for biodegradation is done by gas chromatographic
hydrocarbon fingerprinting which reveals the presence of naphthene biomarkers [6]. Biodegradation does
not generally occur in reservoirs above 80 °C. However, hydrocarbon fluids often have complex histories
over geological time scales. Thus, current reservoir temperature is not always a valid indicator of biodegra-
dation since the hydrocarbons may have migrated. Although both anaerobic and aerobic bacteria can be
active in a reservoir, it is the aerobic bacteria in the presence of a source of migrating or flowing water with
dissolved oxygen that gives rise to the most profound microbial oxidation of crude oil.

Mild biodegradation of crude oil will reduce the fraction of saturated hydrocarbons, and increase the or-
ganic acidic and natural surfactant content. As biodegradation proceeds, progressive loss of n-alkanes is
followed by loss of isoprenoids which leaves an oxidized naphthenic residue. Aromatic functional groups,
resins, and asphaltenes are not attacked and will therefore increase in mole fraction as the other components
are consumed. As discussed below, biodegradation often results in a high concentration of polar organics,
organic acids, natural surfactants, and surface active components which stabilize water in oil and oil in water
dispersions and therefore make water treating difficult.

2.3.3 Hydrocarbons dispersed in water


The treatment of waste water streams from many E & P operations is focused on the removal of dispersed
hydrocarbons. Dispersed hydrocarbons are relatively insoluble in water and are present as discrete hydrocar-
bon droplets within the continuous water phase. Crude oil streams are usually largely composed of aliphatic
hydrocarbons which have a low solubility in water. Thus the hydrocarbons in waste waters from oil production
operations are typically present as dispersed hydrocarbons. One of the critical characteristics of dispersed hy-
drocarbons is the droplet size distribution of the dispersed hydrocarbon phase. The performance of most deoil-
ing equipment will be ultimately limited by the smallest hydrocarbon droplet size that can be efficiently removed
from the water stream. Droplets of dispersed hydrocarbons smaller than the minimum cut off size will not be re-
moved. Droplet size distributions are discussed in more detail in section 2.4. The majority of deoiling equipment
currently installed in E & P operations is only capable of removing dispersed hydrocarbons from waste waters.

2.3.4 Hydrocarbons dissolved in water


All produced water contains dissolved organic compounds. In the past, until roughly ten years ago, produced
water specialists have tended to downplay the importance of these compounds because they are difficult to
characterize. Also, most legislative bodies do not emphasize the importance of dissolved hydrocarbons as
much as they do dispersed hydrocarbons. But it is now recognized that dissolved hydrocarbons in particular,
and organic carbon in general play a major role in flow assurance and in the performance of separation
equipment and that they have a major impact on produced water quality.

By way of example, a class of compounds of major importance are the tetra-acid naphthenates. These
compounds were first identified in the oilfield in the 1960’s. Their importance was demonstrated with the
shutdown of Chevron’s Kuito field in the 1990’s due to calcium naphthenate scaling. The industry responded
with significant activity directed at analysis and understanding of this important class of compounds. Analyti-
cal tests were developed to detect naphthenates, and phase equilibrium models were developed to estimate
their stability. It has since been recognized that naphthenates are not only responsible for well publicized
shutdowns, but they are also responsible for a much more pervasive occurrence of water treating difficul-
ties. Colin Smith, an expert in naphthenates scaling believes that naphthenates solids precipitation occurs in
roughly 10 % of North Sea crude oils, 20 % of west African crude oils, and 30 % of south east Asian crude
oils [7]. But naphthenates are only one of many organic compounds in produced water.

Organic material is considered to be dissolved if it passes through a 0.45 micron filter. Particulate or dis-
persed oil in water is that which is retained on the filter. Obviously this is an arbitrary criterion. Nevertheless,

Deoiling Manual V 2.0 17


Restricted

it is universally adopted in the E&P industry. The 0.45 micron filter is commonly used for many other purposes
and is readily available. One of the problems with this criterion is that drops of light viscosity oil tend to pass
the filter whereas drops of heavier oil do not. Nevertheless, the 0.45 micron filter provides a simple and
readily available test and has been accepted as the industry convention.

2.3.5 Dissolved organic carbon


Once the sample has been filtered, then a simple and commonly used test to measure the concentration of
organic compounds is the dissolved organic carbon (DOC) test. DOC is measured by converting all of the
organic material in solution to CO2 and then measuring the CO2 that is produced. TOC is the organic con-
tent measured using the same test as DOC but run on the unfiltered sample. ASTM D 2579-78 discusses test
methods for measuring total and organic carbon concentration. It references information relating the TOC to
other measures of water quality such as biological oxygen content (BOD), and Chemical Oxygen Demand
(COD), which are both used to characterize produced water effluent into rivers lakes, streams, and other
biologically sensitive areas.

DOC is often used to assess the biological nutrient content of injection water for waterflooding. High DOC
content water will likely result in significant reservoir souring (conversion of sulfates to H2S) if sulfate is present.
Seawater, collected at least a few hundred meters offshore, typically contains about 0.5 to 5 mg/L of dissolved
organic matter, as measured by DOC. Produced water can contain a few mg/L to many hundreds of mg/L.

Identification of dissolved organic compounds in produced water is usually made from two perspectives.
One perspective is to determine the concentration of specific relatively low molecular weight species such as
benzene, toluene, the xylene isomers, ethyl benzene, phenol, benzoic acid, and the Short Chain Fatty Acids
(SCFAs) such as formic acid, acetic acid, propanoic acid, and butanoic acid. In fact, SCFA is a misnomer
since most of the compounds usually included in this list have no or almost no aliphatic chain. Another name
for these compounds is Volatile Organic Acids (VOCs) but the former name is more common. The importance
of all of these compounds is discussed below.

The other perspective that is used together with species identification is to characterize the larger compounds
using functional group analysis or CHNOS (C, H, N, O, S pronounced “cheenose”) analysis. Functional
group analysis usually focuses on the following groups: carboxylic acids, substituted phenols (phenol, o-
cresol, resorcinol, hydroquinone, etc), amines, amino acids, humic compounds, and to lesser extent func-
tional groups such as alcohols, ethers, and esters. In general, functional group analysis focuses priority on
those groups that result in surface activity, transition metal binding, and aqueous solubility in the organic
compound.

Environmental risk assessment studies have shown that the aromatic components in produced water constitute
a major contribution to the Environmental Impact Factor (EIF). Recent studies have shown that alkylated phe-
nols and Polycyclic Aromatic Hydrocarbons (PAH) have finite partitioning into both the oil and water phases,
as shown in the table below.

Table 2. Partitioning of PAH components for two different oil in water concentrations. Partitioning percentages are given in mass percent [8].

10 mg/L 100 mg/L


EIF Group oil water oil water
naphthalenes 57 43 81 19
2 – 3 ring PAH 67 33 92 8
4 – 6 ring PAH 61 39 94 6
C0 – C3 phenols 0.05 99.95 0.5 99.5
C4 – C5 phenols 4 96 29 71
C6 – C9 phenols 64 36 95 5

18 Deoiling Manual V 2.0


Restricted

2.3.6 Cations and anions (including heavy metals)


Once the dispersed oil and suspended solids, and conglomerate, have been removed from a sample by
filtration, the dissolved components can be analyzed. As described previously there are dissolved hydrocar-
bons and dissolved organics that play a significant role in water chemistry and the performance of produced
water treating equipment. Many of those components are dissolved in the form of cations or anions. Those
components have already been discussed above and will not be discussed further here. This section deals
with the remaining anions and cations which include the dissolved minerals, metallic cations, transition metal
cations, and anions such as the halides, carbonates, sulfate, sulfides, etc. The major cations and anions of in-
terest and their importance in produced water characterization is given in table 3. With respect to produced
water treating, most of the cations and anions are of importance because of the tendency to form mineral
precipitates, which can then become partially oil wet and stabilize oil in water emulsions.

2.3.7 Salinity
The salinity of water is typically expressed as the concentration of Total Dissolved Solids (TDS), indicating the
quantity of dissolved inorganic salts present in the water.

The salinity of produced water should be considered for the following reasons:
■■ For surface disposal, the salinity of the discharge water should not be significantly different from the salinity
of the receiving environment. This should encompass both the discharge of saline water into fresh waters
and the discharge of fresh water into saline waters (e.g. inshore).
■■ Salinity affects the density of the water which will in turn influence the design of gravity based separation
equipment and the dispersion characteristics of disposed waste water.
■■ Changes in the composition or process conditions of saline waters (e.g. temperature changes) may lead to
the precipitation of inorganic salts and scale formation.
■■ The salinity and aeration of the water will add to the corrosion potential of the water stream, influencing
the selection of materials for process equipment.

2.3.8 pH
The pH determination of water is a relatively reliable indication of its acidic or alkaline tendency. However it
is not a measure of the actual acidity or alkalinity in a water sample. ASTM D 1293-78 discusses the meas-
urement of pH while ASTM D 1067 discusses the measurement of actual acidity or alkalinity.

The pH of the waste water is subject to regulation in some countries which typically require the pH of dis-
charged waste water to be between 5.5 and 9.0

Alkaline waste water with a pH in the range of 8-10 may react with components of the hydrocarbon stream
to form surfactant type chemicals. These surfactants may assist in stabilizing emulsions.

2.3.9 Hardness
Hardness in water is generally caused by the presence of calcium and magnesium ions, though any polyva-
lent cation can contribute to hardness. Hardness in water can result in scale formation in process equipment.
The release of dissolved CO, during depressurisation may also alter the hardness characteristics of water.

2.3.10 Treatment chemicals


A wide variety of treatment chemicals may be present in produced water streams. Typical examples include
corrosion inhibitor, scale inhibitor, demulsifiers, hydrate inhibitors, biocides, flotation aids etc. Other inciden-
tal chemicals may also be present such as detergents used for wash down purposes.

The effect of such chemicals can be significant in the selection and design of deoiling equipment. Many
treatment chemicals are surface active and may stabilize small hydrocarbon drops in the water phase. The
corrosion inhibitors used in gas production operations and the demulsifiers used to assist oil dehydration if not

Deoiling Manual V 2.0 19


Restricted

properly selected, can sometimes result in such stabilisation. The resulting small drops may be very difficult to
separate with conventional deoiling equipment.

Table 3. Major Anions and Cations and their Significance in Produced Water Treating

Cation Significance
Together with chloride, it forms the major component of total dissolved solids.
Does not usually cause problems, except when NaCl concentration exceeds
roughly 25 wt % and halide solubility should be examined as a possible
Sodium precipitate.
Usually present at much lower concentration than calcium ions. Magnesium
carbonate will co-precipitate with calcium carbonate and therefore increase
the mass of carbonate scale. Magnesium forms ion pairs with the sulfate ion
which decreases the mass of sulfate scales thus increasing the solubility of
Magnesium sulfate minerals such as CaSO4.
Major constituent of oilfield brines. Readily combines with bicarbonate anion
to form CaCO3 precipitates. Solubility of CaCO3 decreases with increasing
temperature and decreases with increasing pH. Solubility must be examined
taking into account temperature, pH, pressure (w.r.t. CO2 equilibria), magne-
Calcium sium, TDS. Discussed further below.
Forms BaSO4 precipitate and scale which is difficult to remove, at low con-
centrations. The solubility of BaSO4 is low but increases with salinity (TDS)
and increases with temperature at relatively low temperature and decreases
Barium with change in temperature at high temperature, depending on TDS.
Forms SrSO4 precipitate and scale. More soluble than BaSO4 but somewhat
Strontium similar in properties.
Found in the ferrous state (Fe+2) or ferric state (Fe+3) in produced waters that
have not been exposed to oxygen. Originates from formation brine or from
pipeline corrosion. Once exposed to oxygen, ferric oxides form which are
very insoluble. Ferric oxides form many small solid particles that easily be-
come oil wet thus contributing to stable emulsions and intractable produced
water problems. Various iron sulfide solid compounds will form in the pres-
ence of H2S which, like iron oxides, form many small particles that are easily
Iron oil wet. Major component of Schmoo.
Anion Significance
Major anion counterpart to sodium and major constituent of TDS. Chloride
Chloride increases the corrosivity of steel, which releases iron into solution.
Major problematic anion due to tendency to form sulfate precipitates (CaSO4,
BaSO4, SrSO4) and as a source of sulfur for anaerobic sulfate reducing bac-
teria which generate H2S. Sulfate is a major constituent of seawater. May or
Sulfate may not be present in oilfield brines.
Can react with calcium, magnesium, iron, barium and strontium to form pre-
cipitates. Present in almost all oilfield brines. At relatively low pH it is the
major form of carbonate resulting from CO2/carbonate equilibria. Major
Bicarbonate constituent of alkalinity.
Similar to bicarbonate in terms of reactivity. At relatively high pH it is the
major form of carbonate resulting from CO2/carbonate equilibria. At neutral
Carbonate to low pH it is rarely found.

20 Deoiling Manual V 2.0


Restricted

Some treatment chemicals may be incompatible with other chemicals. Examples are demulsifier chemicals
interfering with subsequent deoiling chemicals, or deoiling chemicals reacting with dilute polymers present in
the water as a result of enhanced recovery schemes.

Efforts should be made to reduce the use of treatment chemicals whenever possible on the basis of minimis-
ing both operating costs and the ingress of additional chemicals into the environment. Important steps are to
investigate the effect on water quality also when selecting demulsifiers, and to check for compatibility of the
various chemicals considered.

2.3.11 Characterization of solids


Solids in hydrocarbon and produced water streams can originate from the formation, or by precipitation
of inorganic scale forming minerals, or precipitation of organic materials. Typical solids in produced water
systems include formation fines (calcite, sandstone), mineral scales such as carbonate (calcium carbonate,
magnesium carbonate, iron carbonate), sulfate (calcium sulfate, barium sulfate, strontium sulfate), or sulfide
(iron sulfide), corrosion products (hematite and magnetite), salt (halite), mineral/organic combinations (cal-
cium naphthenates), organics (asphaltenes, waxes, soaps), and production chemicals (carbamates, polymer
flocs), etc. As far as oil in water emulsions are concerned, the most important properties of produced water
solids are the size distribution, the overall quantity of solids, and the surface wetting characteristics. The treat-
ment and discharge of drilling cuttings and drilling mud is not discussed in this manual.

A good starting point for the analysis of solids in produced water is the NACE Standard Test Method™ 0173
– 2005. Without going into details, solids are typically sampled using filtration through a 0.45 micron filter.
Several sets of sample are collected so that the samples can be subjected to various analyses which include
gravimetric determination of the fraction of organic material, inorganic material, and the composition of the
organic and inorganic constituents. A typical result is given here for solids collected on the Bullwinkle platform
in the Gulf of Mexico in October 2006. These solids were collected from the produced water discharge of the
wet oil tank. This discharge line was fed into a centrifugal pump and the stream was then recycled into the Bulk
Oil Treater. A solids size analysis was not carried out but based on settling tests, the solids were very small di-
ameter (less than a few microns) and had oil attached making them roughly neutrally buoyant in the produced
water. At the time, the platform had several subsea production systems from significantly different reservoirs and
the Total Dissolved Solids content (TDS) of the combined produced waters was high (above 200,000 mg/L).

Table 4. Solids Analysis for solids collected on the Bullwinkle Platform (Oct 2006).

weight %
Gravimetric Wash Test of dry sample Comments
Deionized water wash 19.7 Includes substances soluble in water such as salts
Includes substances soluble in xylene such as paraffin, oil, and
Xylene wash 9.6 organics
Includes substances soluble in weak acetic acid such as carbon-
Acetic acid wash 27.7 ate mineral scales
Includes substances soluble in 15 % HCl acid such as iron sul-
Hydrochloric acid wash 27.4 fide, and iron oxide.
Includes substances insoluble in 15 % HCl acid such as sulfate
Acid Insolubles 15.6 scale, sand, silica fines
Total 100

A portion of the sample was washed using deionized water followed by xylene and then subjected to inor-
ganic solids composition analysis using EDAX, XRF and XRD. The results are given as:
■■ Positive for calcium carbonate scale
■■ Positive for barium sulfate scale
■■ Positive for silica fines

Deoiling Manual V 2.0 21


Restricted

■■ Positive for iron compounds


As given in the table above, it can be concluded that roughly 70 wt % of the sample is composed of these
inorganic solids. Another portion of the sample was washed with deionized water, followed by xylene, fol-
lowed by HCl. This leaves the Acid Insolubles fraction and this sample was analyzed using EDAX and XRD
which gave a positive indication of sulfate scale and silica fines. The acid wash did not evolve noticeable
levels of sulfur which rules out the presence of iron sulfide compounds. Optical microscopy verified the pres-
ence of cubic crystals which are typical of halide precipitates.

Organic solids analysis was performed using DSC/TGA, GC, and H NMR. The results are positive for
waxes and asphaltenes.

Results such as these must be interpreted by the analytical laboratory since they involve a combination of
qualitative and quantitative techniques. However, the final result in this case turned out to be close to that
given in the table of gravimetric wash results.

Table 5. Final results of solids sample analysis

Substance Weight %
Organic material 10
Halide salt from high salinity 20
Calcium carbonate precipitate 28
Iron oxide solids 27
Barium sulfate precipitate and silica fines 15
Total 100

The conclusions from these analyses and observations were:


■■ mineral solids were being formed from a combination of processes (incompatibility leading to barium sul-
fate precipitation, pressure drop leading to carbonate precipitation);
■■ the solids were a conglomerate of inorganic precipitates and oil making them roughly neutrally buoyant in
the high salinity produced water;
■■ high iron oxide solids content indicated corrosion processes which was verified by the frequent requirement
to repair and replace sections of the produced water piping;
■■ high solids content was a major contributing factor in poor water quality, and was a complicating factor
in implementing clean-up technologies which suffer performance degradation in the presence of solids.
■■ Over a period of several months, the problems were solved by a combination of chemical treatment, better
corrosion prevention, segregation of fluids, re-routing of recycle streams.

From an environmental standpoint, suspended solids may have a number of potential impacts, including:
■■ Some solids may be toxic themselves, contain toxic elements, or have radioactive constituents.
■■ Solids may trap or collect other contaminants (i.e. oil).
■■ Discharged solids may accumulate as mud or silt in the local environment.
■■ Discharged solids may result in turbidity in receiving environments with poor dispersion characteristics.
■■ Suspended solids may have a significant impact on the performance of deoiling equipment. The turbidity
caused by discharge of solids may have an environmental impact as well as being undesirable visual pol-
lution.

2.3.12 Impact of solids on water treating


Many of the solids found in produced water systems, particularly the mineral precipitates and corrosion
products are initially water-wet. In some cases, the solid particles remain water wet and are discharged in
the overboard water stream. Some regions of the world have a limit on the amount of solids that can be
discharged. In other regions there is no limit, however the turbidity caused by discharge of solids may have
an environmental impact as well as being undesirable visual pollution.

22 Deoiling Manual V 2.0


Restricted

The solids may settle in the bottom of separators where they occupy volume and thus reduce the residence
time for process fluids. Such solids can contribute to under deposit corrosion or become suitable habitat for
bacteria. Solids can also cause clogging of the piping and precipitate on the walls of vessels and piping to
form scale which is a major threat to flow assurance.

Often though, a fraction of the particles will become at least partially oil wet by attachment of sticky crude oil
components such as asphaltenes, waxes, naphthenates and naphthenic acids, or by attachment of various
production chemicals including corrosion inhibitors and water clarifiers. A solid particle that is initially water
wet and which becomes partially oil wet will then be surface active. Thus, any amphoteric component that
is itself moderately surface active has the potential for wetting the solid surface and enhancing its emulsion
stabilizing properties.

In produced hydrocarbon fluids, potential wetting agents include organic acids, naphthenic acids, as-
phaltenes, resins, waxes, and other polar compounds. In general, the smaller the solid particles, the tighter
the emulsion (smaller emulsion drops). As a general rule, particles that are predominantly water-wet will
stabilize an oil in water emulsion. Likewise, particles that are predominantly oil-wet will stabilize a water in
oil emulsion.

Partially oil wet solids are referred to as conglomerate. Their properties are different from those of oil drops
or solid particles. The density of a conglomerate particle depends on the mass ratio of oil to solids in each
particle, which can vary significantly. The combination of a light component (oil), plus a heavy component
(solid) will result in a specific gravity decrease for the solid particle. When this occurs, separators and
hydrocyclones will be less effective because there is less density difference to drive the relative movement of
the conglomerate from the water. In addition, a fraction of the conglomerate will be neutrally buoyant in the
produced water which will render all gravity separation devices useless. There is some evidence that solids
related oil and water treating problems are not being recognized as such.

From a water treating standpoint, the presence of oily solids (conglomerate) can be devastating. Once a
solid stabilized emulsion forms, it is particularly difficult to separate the components from each other and to
separate the conglomerate from water, due to the shift in density. The more successful technologies for remov-
ing such conglomerate include flotation, chemical treatment, and filtration. Chemical treatment is discussed
below. Filtration carries huge risks in terms of media usage and cost and must be carefully field tested before
implementation.

Flotation is somewhat effective in separating neutrally buoyant solids stabilized emulsions, through the attach-
ment of bubbles which then create the density difference needed for separation. The gas bubbles stick to the
solids themselves and to the oil-wet surface of the solids and carry them to the top of the water where they
are floated over the spill over weir. This is an effective means to perform the separation of oil and water. It
should be noted that while flotation can be highly efficient, it’s application is generally limited to relatively
low levels of conglomerate (e.g., less than 100 to 300 ppm). However, it leads to problems downstream
if the rejects (oil-wet solids) are recycled through the system. This is described in greater detail below in the
discussion of the Bullwinkle platform.

2.3.13 Solids of iron sulfides


Iron sulfide is one of the most problematic solids in produced water treating. Iron sulfide, like iron oxide is
very insoluble. As is typical of the E&P industry, problematic systems are often given descriptive if not color-
ful names. In the case of solids formed from iron, the designated name is “Schmoo.” In gas fields, Schmoo
causes severe operating problems in compressors, pipelines, stabilizer columns, gathering systems, and stor-
age facilities. According to the Gas Machinery Research Council, Schmoo is the least understood and most
prominent contamination problem in pipelines and gas compression equipment.

There have been many reports on the composition of Schmoo, and its precise definition is the source of ongoing
debate. The composition varies considerably depending on water composition and the use of various produc-

Deoiling Manual V 2.0 23


Restricted

tion chemicals such as corrosion inhibitors. For the purpose of this discussion, Schmoo is classified as an oily
solid conglomerate formed fundamentally from iron solids such as iron sulfide and to lesser extent iron oxides.

It is generally true that when a system changes conditions rapidly, such that a relatively insoluble species
reaches supersaturation rapidly, precipitation of the species will form many small particles rather than fewer
large particles. This is particularly true of the iron compounds. Solids formed from iron tend to have size dis-
tributions that average less than a micron in diameter, and their numbers are relatively enormous.

When iron solids are a suspected problem, it is particularly important to practice good sampling technique.
If the produced fluids have relatively high CO2 content (> 1 mole %), then loss of CO2 upon sampling must
be avoided. Loss of CO2 will raise the pH and cause an initially clear sample to turn black due to precipi-
tation of iron sulfide at the higher pH. In that case, iron sulfide may be stable at the lower pH of the in-situ
fluid. Further, if oxygen is allowed to enter the sample then precipitates of iron oxides will form, again where
they may not have been present in the process fluids under process conditions. Release of CO2 or intrusion
of oxygen will also shift the dissolved CO2/bicarbonate equilibria which will affect subsequent analysis of
carbonate stability.

In addition, solids formed from iron are easily oil wet thus making them amphoteric (oil and water loving)
and therefore driving them to the oil/water interface. Further, various surface active compounds (acids, as-
phaltenes, corrosion inhibitors) also tend to bind to the surface of iron solids thus making them attract oil to an
even greater degree. These properties of iron solids have been exploited in the cleanup of produced water
systems through the use of carbamate chemicals which are described further below. However, from what has
been said thus far regarding iron solids, it will be clear that the use of carbamates must be undertaken with
care. Given their small size, large numbers, and wetting properties, solids formed from iron are extremely
effective at stabilizing oil in water emulsions.

2.3.14 Responding to solids problems


It is beyond the scope of this section to provide a complete discussion of how to address solids problems.
However some initial guidance is provided here.

Generally the most effective strategy to deal with solids problems in a water treating system is to prevent
solids production in the first place. Prevention of scale forming mineral precipitates is prudent from a flow
assurance standpoint. In one example, a scale inhibitor added upstream to prevent barium sulfate precipita-
tion had a very beneficial effect on water quality by preventing solids from forming. The Waterflood Manual
[9] discusses scale formation and compatibility of different produced water streams. The Shell Scale Manual
[10] discusses the use of Scalechem modeling package and interpretation of results. Also, there are books
on water treating that cover the subject well [3, 4].

Chemical treatment of conglomerates has been effective in many cases but chemicals for this purpose must be
applied with care. The injection of acid is intended to displace the oil from the solid particle surface and thus
return the solid particles to a water-wet state which then facilitates gravity separation and eventual discharge.
However, acid will almost certainly cause corrosion problems.

Flocculating agents are also used with success to treat solids. When a floc forms, the larger diameter of the
floc magnifies the effect of any density difference with water and facilitates separation by settling or flotation.
Flocculating agents must be applied with care. Flocculating agents must always be applied at a location
where minimal shearing will occur downstream. Shearing does not improve mixing of the chemical. Instead
it will break any floc that has formed and irreversibly render the chemical useless. If the system is over-treated
with a flocculating agent, a sticky polymer floc can form from the combination of oil, solids and polymer.
This combination becomes particularly troublesome in recycle streams. Flocs rejected back into the process
stream can accumulate more solids and oil. This can result in more emulsion, thicker pads in the separator
vessels, and more viscous pads which in turn can lead to fouling of vessel internals, including plugging of
inlet headers, collection headers, distribution screens, treater grids, and level detectors.

24 Deoiling Manual V 2.0


Restricted

2.3.15 Examples of water treating problems caused by solids


The fluids in the deepwater GoM are particularly suitable to forming these conglomerate stabilized emulsions.
They have relatively high asphaltene content from a few to several percent. Resins are also somewhat high
varying in the range of a few percent to 15 % resin. Due to relatively high aromaticity and resin content the
asphaltenes tend to be stable to marginally stable, with none of the fluids in the unstable region. While this
means that asphaltene precipitation is not a problem in the reservoir, near well bore or tubing locations, as-
phaltenes do precipitate to some extent in the topsides and contribute to the stability of both water in oil and
oil in water emulsions. They also stick to solids particles causing solids stabilized emulsions. This, together
with the relatively short residence times found in the deepwater, leads to a relatively challenging situation
regarding oil and water separation. As already discussed, sophisticated electrostatic treater design and
well managed chemical programs are required to separate drops of water from oil that are stabilized by
asphaltene particles. Thus, it is not surprising that flotation would be of such great use.

As a specific example of processes that can lead to water treating problems due to solids, an example from
Shell Brazil is considered. When a reservoir is waterflooded with any water other than source water, there is
the potential for scale formation as the two water chemistries intermingle. In a typical seawater breakthrough
scenario, barium sulfate particles will form in the produced fluids. The particles can be very small (<1 mi-
cron to 2 micron), particularly in the presence of a scale inhibitor. As the barium sulfate particles form, the
asphaltenes may also be precipitating. As asphaltenes precipitate, they may adsorb onto available surfaces
including the barium sulfate particles. Such fine particles, with an asphaltenic coating are excellent water-in-
oil emulsifying agents.

2.3.16 Dissolved oxygen


Produced water does not normally contain any dissolved oxygen. In fact, most produced water has never
been in contact with oxygen. Thus, the dissolved and suspended components in produced water are typi-
cally in a reduced state, in the sense of oxidation/reduction potential. Once produced water is exposed to
oxygen its nature changes dramatically and for the worse in several regards. This fact cannot be overstated.

Introduction of oxygen laden water into the process stream is extremely detrimental because it (1) oxidizes the
iron in solution in the water, creating iron oxide solids and (2) creates very aggressive, damaging corrosion in
the form of pitting caused by the formation of anodic sites on internal metal surfaces. Iron in the fluids, when
oxidized, will precipitate and contribute strongly to the stabilization of the emulsions.

Bacterial contamination becomes a greater problem in the presence of oxygen since aerobic bacteria are
between 10 to 1,000 times more active than anaerobic bacteria. Aerobic bacteria grow significantly faster,
and multiply significantly more frequently than anaerobic bacteria. Bacterial activity results in corrosion, and
generation of sticky biopolymers which contribute to the stability of oil-in-water emulsions and which can
contribute to pad formation in separator vessels. A few common sources of oxygenated water entering into
a separation system are by way of a seawater sump system, holding tanks for off-spec produced water, or
an open drain system. Such water systems should always be segregated from the production streams. Also,
rigorous biological control should be practiced in such systems.

Another common source of oxygenation of produced water is by way of an open API separator system. Such
systems are very cost effective and have been used successfully in the industry. However, they are open to
the atmosphere and as such must have a means of mechanically collecting the scum that will form on the
surface and the sludge that will form on the bottom of the separator. These accumulations will contain solids
and sometimes sand and dirt that is blow into the separator, and a separate waste treatment facility must
be employed for ultimate treating. Most importantly, any of the reject streams from an API separator (scum,
sludge and oily reject) should absolutely never be routed back into any upstream section of the oil/water
separation train or water treating system.

Deoiling Manual V 2.0 25


Restricted

2.4 Droplet size distributions


The droplet size distribution is one of the most crucial characteristics governing the removal of oil from water.
Many designs of deoiling equipment have a limit on the smallest hydrocarbon droplet size that can be ef-
ficiently removed from the water stream.

Figure 2.1 illustrates a typical droplet size distribution curve in terms of hydrocarbon volume. The size distribu-
tion is typically categorised by a mean droplet size, however the mean droplet size must always be defined
on an associated basis, e.g. numerical mean, volume/weight mean, surface area mean etc. When reporting
mean droplet sizes the basis for the mean should always be clearly stated.

A recommended method for expressing mean droplet sizes is the mean volume droplet size Dv,50. This is
defined as the droplet diameter such that half the volume of hydrocarbons present is encompassed in smaller
droplets and half the volume is contained in larger droplets. Graphically, as illustrated in Figure 2.1, the
mean volume droplet diameter divides the volume distribution such that half the area under the curve is on
either side of the mean, in this case the mean volume diameter being 26.6 µm.

The volume based mean droplet diameter can be more readily determined by plotting the cumulative percent-
age volume as indicated in Figure 2.2.

The numerical mean, defining the point where half the number of droplets are either larger or smaller is often
not as useful for indicating the distribution of dispersed hydrocarbons. The volume of hydrocarbon contained
in the smaller droplets may be insignificant when compared to the larger droplets. For example, a single
50 µm droplet contains 1000 times the volume of a 5 µm droplet.

The numerical mean for the distribution illustrated in Figure 2.1 was separately determined to be 2.3 µm,
significantly lower than the volume based mean. Thus, even though half the number of droplets are smaller
than 2.3 µm, from Figure 2.2 we can see that volume of hydrocarbons contained in these droplets is less
than 5% of the total hydrocarbon droplet volume.

26 Deoiling Manual V 2.0


Restricted

2.5 Characterization checklist and ranking


The following table summarizes the characteristics that will likely lead to water treating problems.
■■ Solids (in general) and Iron Sulfides (in particular):
−− Oil wet solids (conglomerates)
−− Neutrally buoyant conglomerates
−− Small particle size and/or high concentration of solids
■■ Biodegradation:
−− High TAN or high concentration of organic acids
−− Presence of calcium naphthenate
■■ Unstable asphaltenes:
−− High ratio of asphaltene/resin concentration together with high saturates/aromatics
■■ Production Chemicals:
−− Presence of methanol, anti-agglomerants, hydrate inhibitors, corrosion inhibitors
−− In particular over-dosing of these chemicals

Table 6 gives a checklist of the main parameters that should be measured in a characterization of produced
water. An assessment can be made of the difficulty of oil/water separation on the basis of a point system. For
any water characterization, the total ranking points should be added. The second table below then provides
a ranking which can further be used to determine the difficulty of the separation problem.

Deoiling Manual V 2.0 27


Restricted

Table 6. Quantitative Assessment of Water Characteristics

Characteristic Source Design Detail Possible Characteristic Ranking


Points Points
Stokes Factor API, temperature, Longer residence 10 < 200,000 10
water density time in primary 400,000 to 800,000 5
separators
> 800,000 0
Oil flow assur- Wax, paraffin, Inhibitors, heat- 5 No inhibitors or heating 2
ance factors asphaltene stabil- ing for wax and required
ity, incompatible paraffin Inhibitor or heating required 5
hydrocarbons and none used
Inhibitor or heating required 0
and is used
Biodegradation TAN, fingerprint- Secondary separa- 15 0 wells 0
ing, biomarkers tion equipment, 1 to 2 wells 10
optimized chemi- Delta API > 2
cal treatment
> 2 wells 15
Delta API > 2
Scaling ten- Mineral scales Water wetting 5 No inhibitors required, or in- 0
dency e.g. carbonates, chemicals, filtration hibitor required and is used
sulfates; incom- / tertiary separa- Inhibitor required and none 5
patible water tion equipment used
Dissolved or- Acids, naphthe- Secondary separa- 5 pH > 6, acid < 100 mg/L 0
ganics nates tion equipment, 4 < pH < 6 3
optimized chemi- 100 < acid < 500
cal treatment
pH < 5 5
acid > 500 mg/L
Solids Formation fines, Water wetting 15 Solids < 100 lb/MBbl 0
scale particles, chemicals, filtra- 100 < solids < 400 10
tion/tertiary sepa-
Solids > 400 lb/MBbl 15
ration equipment
Iron sulfide H2S, iron Secondary separa- 20 FeSx < 10 mg/L 15
tion equipment, 10 < FeSx < 50 18
optimized chemi-
FeSx > 50 mg/L 20
cal treatment
Surface ac- Corrosion inhibi- Secondary separa- 10 No change upon turning off 0
tive or shear tor, tion equipment, chemical
enhancing methanol optimized chemi- Moderate water deteriora- 5
chemicals cal treatment tion w/chemical
Severe water quality deterio- 10
ration w/ chemical
Small drops High shear Reduce shearing 15 D < 10 micron 15
10 < D < 50 10
D > 50 micron 0

28 Deoiling Manual V 2.0


Restricted

Severity of treating challenge:

Type 1 System:
Total points less than 35
No iron sulfide.
No biodegradation.
For this level of separation challenge, a typical system may consist of primary separation followed by hydro-
cyclones. Flotation may be required, depending on the Stokes Factor. See the example below.

Type 2a System:
Total points between 35 and 55
No iron sulfide.
No biodegradation.
For this level of separation challenge, a typical system may consist of primary separation, hydrocyclones and
flotation. Care should be given to the handling of reject streams from the water treating equipment in order
to ensure that a stabilized emulsion is not generated.

Type 2b System:
Total points between 35 and 55
Presence of iron sulfide or biodegradation.
For this level of separation challenge, a typical system may consist of primary separation, hydrocyclones,
flotation and some means of treating the reject from the water treating equipment. Chemical application will
be critical both in terms of demulsifier and deoiler selection and optimization, but also in terms of minimizing
the use of methanol and corrosion inhibitor.

Type 3 System:
Total points above 55
Presence of iron sulfide and/or biodegradation.
For this level of separation challenge, a typical system may consist of primary separation, hydrocyclones,
flotation and some means of treating the reject from the water treating equipment. Chemical application will
be critical both in terms of demulsifier and deoiler selection and optimization, but also in terms of minimizing
the use of methanol and corrosion inhibitor. Some form of tertiary water treating equipment will be required
such as filtration.

2.6 Examples of produced water characterization


In this section we give examples where the results of produced water characterization were used in the de-
sign of a new field or to solve a water treating problem. In the examples given here, only the highlights of
the characterization are given, rather than the complete characterization.

2.6.1 North Sea versus Mars Deepwater Gulf of Mexico


As discussed above, one of the simplest characterizations that can be carried out for fluids is to calculate the
Stokes Factor. A Consultancy Services Benchmarking report gives Stokes Factors for BP and Shell platforms
in the North Sea. Below data are provided from which the Stokes Factor can be calculated for the Mars
platform in the deepwater Gulf of Mexico. The Stokes Factor is a measure of the settling tendency of an oil/
water fluid, without taking into account the drop size. The higher the Stokes Factor, the greater the speed with
which oil drops rise to the surface of the water phase in the separator vessel.

Mars fluids:
■■ API 17.3 to 31.5
■■ Water density: 1,140 kg/m3
■■ Water temperature: 120 °F (50 °C)
■■ Water viscosity: 0.56 cP

Deoiling Manual V 2.0 29


Restricted

■■ Stokes factor 337,605 to 485,539

Also provided in the Consultancy Services report is the volumetric flux of liquids through the main oil/water
separator. The volumetric flux is a measure of the flow related duty of the separator. It is calculated as the
ratio of the total liquid volumetric flux and the interfacial area between the oil and water in the separator. As
the flux increases, separation becomes less efficient.

Mars FWKO:
■■ Total liquid volumetric flow: 200,000 BFPD = 0.368 m3/sec
■■ FWKO: S/S length: 40 feet; Inside diameter: 120 inch; NWL from bottom: 3.3 feet
■■ Flux rate: 0.0090 m3/sec/m2

Data for both Shell and BP in the North Sea, together with the Mars value for the deepwater Gulf of Mexico
are given in the figure below. As shown, the flux rate is significantly higher for Mars which means that vessel
residence time is lower. It is noted that in typical North Sea platforms, water treating includes hydrocyclones
but not flotation. In the deepwater GoM, both hydrocyclones and flotation units are required to meet very
similar specifications.
0.014
Shell North Sea
0.012 BP North Sea

0.010 Shell Deepwater GoM


Flux Rate (m3/(s m2))

0.008

0.006

0.004

0.002

0.000
0 200,000 400,000 600,000 800,000 1,000,000 1,200,000
Stokes Factor (kg/(m Pa s)) 3

Figure 2.3. Flux Rate in the main oil/water separator versus the Stokes Factor which provides a simple means of characterizing the fluid properties.

2.6.2 Deepwater Gulf of Mexico example


At the time of this writing, Shell operated six deepwater platforms in the Gulf of Mexico (Bullwinkle, Auger,
Mars, Ram Powell, Ursa, Brutus). The platform with the most problems at the time was Bullwinkle. In the fol-
lowing the fluids characteristics of Bullwinkle will be compared to those of the other platforms. Fluid similari-
ties between Mars and Bullwinkle suggested that techniques used on Mars would have a good chance of
working on Bullwinkle as well.

The characteristics of the reservoir sands, and of the hydrocarbons are included as well. The properties of
produced water are strongly affected by both.

As shown in Table 7, the fluids range in °API from 24 to 32, with most wells producing in the range of 27
degrees API. The Ram-Powell field, and the early Auger field are exceptions. Both had relatively high gas
production with gas condensate liquids. The gravity of the condensate was in the range of 32 to 34 °API.
Over time, Auger reservoir targets moved from gas to oil.

30 Deoiling Manual V 2.0


Restricted

Most of the hydrocarbon fluids produced at the Shell deepwater locations are of the Miocene age. The reser-
voirs are typically deep (deeper than 15,000 feet below subsurface). Reservoir temperatures are in the range
of 75 to 80 °C. All of these factors suggest that biodegradation is possible but not highly probable and is
likely determined by access to percolating, low saline and nutrient rich water. As previously mentioned, given
the low occurrence of aquifer volumes, and the fact that there is a high occurrence of massive salt bodies,
there is little evidence that access to such percolating low saline water has occurred. The fluid properties do
indeed reflect this analysis.

While biodegradation is relatively rare across the Shell deepwater GoM portfolio, it was seen in two notable
exceptions. The Pink reservoir on Mars and Bullwinkle are both considered to be biodegraded. For Mars
Pink, the acid number is high (see table below), the relative fraction of saturates is lower and the aromatic
fraction is higher. GC fingerprint analysis also showed characteristics of biodegradation. Platform personnel
noticed that when the well containing biodegraded fluids (A-1) were flowing, oil dehydration (< 1% BS&W)
and water deoiling (< 29 ppm) targets were much more difficult to achieve. Biodegradation was also de-
tected in the Bullwinkle Pink reservoir which also had high Total Acid Number and caused water treating
problems when online. Also, naphthenic acid content for the platform as a whole was rather high and this
was thought to be due to the wells flowing from the Pink reservoir. Across the other Shell deepwater platforms,
it is thought that the biodegradation ceiling is below the Pink reservoir but above all the other reservoirs.

Biodegraded fluids are problematic because they contain a relatively high concentration of surface active
species such as short chain fatty acids, acidic resins and acidic polynuclear compounds. Mars not only had
such fluids to contend with but it also had short residence times as well. The water treating system was typical
of GoM deepwater consisting of a FWKO, hydrocyclones, a Wemco flotation unit with recycle to a slop
tank and ultimately to the Bulk Oil Treater. To handle the biodegraded Pink fluids on Mars, a rapid acting
high molecular weight polyol type demulsifier was selected from bottle testing. This demulsifier was somewhat
expensive but it helped maintain good control of BS&W without degrading produced water quality. Vessels
were kept clean of pads. A carbamate water clarifier was used. More details can be found in [11].

While wax is not a general problem at Shell deepwater locations, there are a few wells that require wax
inhibitor or some form of thermal or mechanical treatment to remove wax buildup. For the most part, wax
problems are isolated to a few wells.

All of the Shell deepwater GoM production is sweet with no H2S. There were no signs of iron sulfide or
Schmoo. The CO2 content generally falls in the range of 0.1 to 0.4 mole % by volume for gas recovered
from a multiphase sample at stock tank conditions. This is relatively low and helped to maintain relatively low
bicarbonate concentrations. Although the produced water was highly saline, and contained high concentra-
tions of calcium and magnesium, calcium carbonate scaling was not a major problem, due to relatively cool
temperatures and low bicarbonate concentrations.

As a detailed example of fluid property variation within a field, fluid properties for the Mars reservoirs are
given in the table below. The gravity of these fluids is included in Table 7. The asphaltene and resin content
are also shown in Table 7.

Deoiling Manual V 2.0 31


Restricted

Table 7. Fluid Properties for the Mars Reservoirs

Atomic S ACID
Reservoir DEPTH API Saturates Aromatics Resins Asphaltenes (wt %) No.
Pink 13,036 17.3 21.0 61.6 14.0 3.4 2.7 4.35
Lower Green 16,287 31.5 40.9 47.6 9.5 2.1 1.8 0.34
Ultra Blue 16,301 24.6 41.3 37.5 13.9 7.3 2.6
Orange 16,550 24.4 27.4 54.5 14.5 3.6 2.4 1.00
Upper Green 16,910 23.5 24.3 55.6 14.9 5.2 2.6
Magenta 17,610 24.2 28.1 54.1 13.5 4.3 2.6 0.61
Violet Ic 18,419 22.5 25.0 59.3 12.6 3.1 2.7 1.20
Lower Yellow 18,476 27.6 30.8 51.8 15.1 2.2 2.2
Terra Cotta 18,476 22.1 25.2 50.2 13.4 11.2 2.8 0.95

Note that atomic sulfur in the Mars crude is mostly incorporated in aromatic species such as thiophenes,
benzothiophenes, dibenzothiophenes, and higher molecular weight species.

Most of the Shell deepwater reservoirs are bounded by large salt bodies. What little aquifer water there is
in these basins is typically in communication with these salt bodies along their outer edges. Thus, as shown
in Table 7, and as the typical analysis shows in Table 8, dissolved mineral content of the produced water
tends to be very high.

Table 8. A typical Shell deepwater GoM water analysis – Mars Terra Cotta A-4 well

Specific Gravity 1.13


pH 6.5
Cations mg/L
Sodium 61,300
Calcium 5400
Magnesium 1630
Barium 160
Iron 20
Anions mg/L
Chloride 109,000
Bicarbonate 120
Sulfate 1
Total Dissolved Solids (TDS) 177,640

The incidence of solids in the produced fluids in the GoM has only been sparsely studied. Some results are
given in Figure 2.4. The solids found in the GoM are small diameter (3 micron and smaller) sand stone fines.
On Mars (GoM-1 in the Figure) and Bullwinkle (GoM-2 in the figure) there was a tendency for these fines to
be coated with asphaltenes which contributed to produced water problems.
Once a solid stabilized emulsion forms, it is particularly difficult to separate the components. A fraction of
the emulsion drops, if dispersed into the water phase, will be neutrally buoyant in the produced water. That
is, depending on the relative quantities of oil and solids, the combination of a lighter than water component
(oil), plus a heavier than water component (solid) can result in a specific gravity close to that of water. When
this occurs, separators and hydrocyclones will not provide effective separation because there is insufficient
density difference to drive the relative movement of the oil and water. Solids related oil and water treating
problems are often not being recognized as such.

32 Deoiling Manual V 2.0


Restricted

1200

1000
lbs / MBbl (ASTM D-4807)

800

600

400

200

ira 1

na
ns a

be a
f o off la
ex re

M -3
C ern exi g

M -2
rn ne 1

lb - 2
W o yo 1

be -2
n ll
ay a ll
Su uel

Al tan
Su we
id ale we
n
W et -

-
G ia - zue
f M sho
ifo Ve o -

ta
es f M mi

Al set
jp rta

rta
rta
o
ez

on
on
c

ic
e

m
n

up
Ve

A
ay

S
n

Bi
er

w
f
st

t
ul

ul
id
Ea

al
M

M
ia

ia
rn

rn
ifo

ifo
al

al
C

Figure 2.4. Solids content of produced fluids from various regions of the world. It is important to compare one region to another in order to gain a perspec-
tive on the relative magnitude of solids problems. California Midway Sunset-2, Alberta, and Montana have some of the most challenging solids problems
in the world. Fields such as California Midway Sunset-1 do not have significant solids problems.

In 2001, drop size distribution was measured on the Mars platform. The volume average drop diameter is
given for various locations in the figure below. Unfortunately, not all locations could be sampled during this
study. Particular emphasis was placed on locations that would elucidate the effect of pump shear. Recycle
systems (shown in blue) often require pumping in order to push the fluid into an upstream location. As shown,
the discharge from the two produced water centrifugal pumps did in fact have very small average drop
diameters. A sample of fluid was taken upstream of the Bulk Oil Heat Exchanger. As observed under the
microscope, the fluid was composed of a continuous oil phase in which drops of water were dispersed. In
addition, the drops of water had very small drops of oil dispersed in them. This is typically referred to as a
complex emulsion. The cause of the complex emulsion was the recycling loop shown in the process sche-
matic. Another feature of the data given in the figure is the relatively small drop diameters that were observed
on the platform. In particular, the overboard discharge stream, from the Wemco Induced Gas Flotation Unit
had an average drop diameter of 6 microns. This is considerably smaller than would be expected without
the use of chemicals. But in fact, a fast acting and highly active carbamate chemical was being used. This
could account for the high performance of the IGFU.
A summary of the water treating experience for the deepwater platforms in the time period from 2001 to
2006 is given in the table below. This table identifies the characterization area and gives the characteriza-
tion result together with the observations of the water treating performance and strategies used to overcome
the fluid properties that lead to water treating challenges.

Deoiling Manual V 2.0 33


Restricted

Complex emulsion
Gas Gas
Gas
HEx
Primary Oil
TdeG
to DOT

FWKO BOT

30 mg/L
OiW
100 mg/L OiW Centrifugal pumps
3 micron
30 micron D50 2 x 292 gpm ea
D50

20 mg/L OiW
DVA or 6 micron D50
Hydrocyclone Flotation
Subsea wells

Sump
System

WOT
Oil
340 mg/L OiW to DOT
1 micron D50

Oily water Overboard


Discharge
Figure 2.5. Process Schematic of the Mars TLP oil and water treating system. Volume average drop size is shown as measured by isobaric sampling and
optical microscopy determination of drop size distribution.

34 Deoiling Manual V 2.0


Restricted

Table 9. Correlation between produced fluids characterization and the performance of water treating equipment.

Characterization Learning from Equipment Performance


Subsurface setting (depth, T, geo- High salinity seen across the region which gave high density dif-
chemistry) ferences between oil and produced water; some scale problems
addressed with scale inhibitor; separate trains required in some
cases to segregate incompatible fluids.
Possibility of biodegradation Hand full of wells caused significant oil dehydration and water
deoiling problems when they were producing; extensive vessel
internals modifications were required for oil dehydration; process
re-routing and chemical treatment work required for water deoiling
in order to stay on specification when those wells were producing.
Gas constituents (CO2, H2S) & pro- Sweet production thus no iron sulfide problems; carbonate was
cess conditions (T, P) high in some cases but with low CO2 (< 1 mole %), moderate
bicarbonate, and no need for alkaline H2S scavengers, calcium or
magnesium carbonate was kept under control with scale inhibitors,
and scale monitoring.
Oil characterization Resins and asphaltenes were high; asphaltenes were moderately
stable due to high resins and aromatics, with a few exceptions;
caused foaming, emulsion, and water treating problems. Had to
apply special chemistries, keep water treating equipment clean by
flushing with solvents, developed special acid flow back proce-
dures to prevent acid precipitation of asphaltenes, changed pro-
cess routing to reduce complex emulsions, minimize recycles, avoid
condensate mixing w/ oily water, minimize shearing, improve
treatment of recycle streams.
Brine analysis by asset & by region High iron was used to advantage w/DTC chemistry.
Possible solids from reservoir Cleaned vessels frequently; practiced rigorous sand surveillance;
Auger applied an acid treatment to strip oil off of solids. Solids
wetting demulsifiers applied at Mars.

Deoiling Manual V 2.0 35


Restricted

2.7 References to Chapter 2


1. M.B. Flannery, et al. “Guidelines for Sampling and Analysis of Reservoir and Production Fluids,” EP
2007-3186, SIEP Houston (2008)
2. “Standard Methods for the Examination of Water & Wastewater,” 21st Ed., Managing Editor Mary Ann
H. Franson (2005)
3. C.C. Patton, “Oilfield Water Systems,” Campbell Petroleum Series, Norman, OK (1974).
4. A.G. Ostroff, “Introduction to Oilfield Water Technology,” National Association of Corrosion Engineers,
Houston, TX (1979).
5. Erik Tegelaar, “New Guidelines for the Modified IP 143 / 57 Method,” Baseline Resolution, Inc. (1999).
6. L.M. Wenger, C.L. Davis, G.H. Isaksen, “Multiple Controls on Petroleum Biodegradation and Impact on
Oil Quality,” SPE 71450 (2001).
7. “Naphthenate Deposits, Emulsions Highlighted in Technology Workshop,” Journal of Petroleum Technol-
ogy (July 2008).
8. M.A. Reinsel, J.J. Borkowski, J.T. Sears, “Partition Coefficients for Acetic, Propionic, and Butyric Acids in
a Crude Oil / Water System” J. Chem. Eng. Data v. 39, p. 513 – 516 (1994).
9. Shell Waterflood Manual Chapter 2, “Water Sourcing and Chemistry,” Shell Waterflood Improvement
Team (2008).
10. D.M. Frigo, “Scaling Manual: Inhibition of Oilfield Scales,” SIEP 99-5679 (1999).
11. J. Walsh, T. Frankiewicz, “Treating Produced Water on Deepwater Platforms: Developing Effective Prac-
tices Based Upon Lessons Learned” SPE 134505 (2010).

36 Deoiling Manual V 2.0


Restricted

Appendix 2.1

Stokes’ law

Deoiling Manual V 2.0 37


Restricted

Appendix 2.1 Stokes’ law

Provided a density difference exists between the hydrocarbon phase and the water phase, a dispersed hy-
drocarbon droplet will experience a buoyant force as expressed by equation A2.1.1:

 (A2.1.1)

where: F, = buoyancy force (N), g = gravitational acceleration, (m/s2), d = dispersed droplet size (m),
rw = density of the continuous water phase (kg/m3), rh = density of the dispersed hydrocarbon phase
(kg/m3).

In most cases the hydrocarbon density will be lower than that of the water phase and the dispersed droplet
will experience a positive buoyant force causing the droplet to rise upwards. As the droplet rises through
the continuous phase, the buoyant force will be opposed by a drag force on the droplet. The drag force is
expressed by equation A2.1.2:

 (A2.1.2)

where: Fd = drag force (N), CD= drag coefficient and v = droplet velocity relative to the continuous phase
(m/s).

When the buoyancy force and drag force are in balance the droplet will rise with a constant terminal velocity
vT. For a dilute dispersion of small hydrocarbon droplets in a water phase the Reynolds number is generally
less than 10 and the following approximation for the drag coefficient may be used:

 (A2.1.3)

where: / and µ = dynamic viscosity (Pa.s or cP/103)

Equating the equations for the buoyancy and drag force (A2.1.1 and A2.1.2), and substituting the approxi-
mation for the drag coefficient (A2.1.3) and Reynolds number results in Stokes’ law:

 (A2.1.4)

where: vT = terminal droplet rising velocity relative to the continuous phase (m/s).

Fig. A2.1.1. Forces acting on an oil droplet in water

38 Deoiling Manual V 2.0


Restricted

Chapter 3.

Oil in water dispersions

Deoiling Manual V 2.0 39


Restricted

3. Oil in water dispersions


3.1. General
Water collected from the primary separation process is normally a dilute oil in water dispersion with a few
hundred up to a few thousand ppm oil. This means that from the perspective of the oil droplets the water
phase is an infinite medium and other oil droplets are still far away such that hindered settling and coales-
cence do not occur. The physical properties of the dispersion as a whole are determined by the water proper-
ties, and the behavior of the droplets can be described by Stokes settling (see Appendix A2.1).
For the design of separation equipment it is useful to be able to estimate the size distribution of the oil drop-
lets. This makes it possible, to establish the removal efficiency by comparing the size distribution to the cut-off
droplet size of the equipment.
The droplet size distribution in its turn is determined by the energy dissipation or mixing intensity prevailing
in the upstream handling. In some cases this can be quantified too.
Parts of the production process with a high mixing intensity are:

■■ The near well bore formation


■■ Down hole perforations
■■ Gravel packs
■■ Subsurface pumps
■■ High GOR tubing flow including gas lift
■■ Chokes
■■ Manifolds
■■ Pumps
■■ Control valves

Typically, the higher the rate of pressure loss with time the higher the mixing intensity and hence the smaller
the droplets. Control valves, where the fluid undergoes a large pressure drop almost instantaneously, can
generate very fine dispersions. The actual droplet size distribution that results depends on a number of other
factors including properties of the fluid.

3.2. Mixing intensity and maximum droplet size.


Even when the overall shear intensity is high, there is a distribution of shear rates in the fluid. In a localized
high shear zone, drops will break apart. In a localized zone of moderate or low shear, drops will not break
apart and may actually coalesce. Whether or not the shear rate is sufficient to break the drops depends on
the interaction between two stresses. One of these stresses is external to the drop and is due to turbulence.
The external stress deforms the drop and attempts to break it up. The other stress is internal to the drop. The
internal stress acts to restore the spherical shape of the drop and keep it together. If the shear rate across the
drop due to turbulence is sufficiently high to overcome the restorative stress, the drop will break apart.

The Weber number is a characteristic dimensionless number that gives a measure of the relative magnitude
of these two stresses.

external turbulent stress


We =
internal restorative stress

In turbulent flow the external stress is determined by the inertial pressure drop across the diameter of the drop.
The latter is related to the turbulent velocity fluctuations around the droplet. The external stress is given by:

The internal restoring stress is the interfacial stress given by 2σ/d, where σ is the interfacial tension.

40 Deoiling Manual V 2.0


Restricted

To describe the stability of a droplet the dimensionless Weber number is used, which is the ratio between the
deforming and the restoring stress, In our case, this leads to:

In turbulent flow it can be shown that the velocity fluctuations are related to the energy dissipation rate per
unit mass. This establishes the following relation:

(∆u) = (ε d) 3
1
2

In addition to drop breakup, coalescence can occur in a turbulent fluid as well. If the drop does not break apart,
as in a localized low shear zone, the net effect of turbulence is to increase the frequency of drop to drop colli-
sions which may lead to coalescence, depending on the trajectories and collision rate. The net effect of this shear
rate distribution is to cause some drops to break and others to coalesce with one another. Thus, a dispersion that
undergoes shear experiences both breaking and coalescing processes simultaneously, albeit to different degrees
depending on the shear intensity and flow configuration. Thus, for a given shear field, there will be a maximum
droplet size which is related to the dynamic equilibrium between the forces of drop breaking and coalescence.
What is meant by maximum drop size needs to be defined since the drop size distribution is a statistical quantity
and there may be a finite probability for very large drops. Typically, the maximum drop size is defined as the
95% cut off of the drop size distribution, meaning that 95% of the total oil volume is contained in droplets up to
this size. This 95% cut off corresponds to an empirically determined maximum Weber number (Wecrit).

Substituting the previous expressions and rearranging gives an expression for the maximum drop size in a
turbulent flow:

(3.1)

This expression was developed by Hinze[1]. The value of Wecrit is ≈ 0.725.

3.3. Mixing intensity and pressure drop.


As discussed above, the energy dissipation rate per unit mass is related to turbulent velocity fluctuations. This
is not very practical to work with. An alternative relation, in terms of permanent pressure drop is possible.
Since pressure drop is an easily measured quantity this is far more useful. In an orifice, choke valve, control
valve, or any restriction, the fluid is forced through a relatively small flow volume and therefore the velocity
increases. This results in a decrease in pressure. In this zone, velocity gradients are present which can result in
turbulence with corresponding energy dissipation. Once the fluid passes out of the restricted zone, the veloc-
ity decreases and the pressure recovers. However, the presence of the turbulent zone and its corresponding
energy dissipation, causes a permanent pressure drop compared to the pressure upstream of the restriction.
The pressure drop, which is easily measured, can be related to the energy dissipation rate and hence to a
quantitative measure of shear forces which result in net breakup of oil drops.

The average energy dissipation rate per unit mass of fluid can be derived using the following approximation.
Consider a fluid element. The irreversible pressure drop across this element is denoted by ΔPperm. The energy
loss associated with this pressure loss is given by ΔPperm A Δx, where A is the cross sectional area, and Δx is
the length of the zone of turbulence. The time in which this energy is dissipated is denoted as Δt. The mass of
the fluid element is equal to ρ A Δx. Combining these terms results in the following expression for the average
energy dissipation rate per unit mass of fluid:
(3.2)

Deoiling Manual V 2.0 41


Restricted

Frominspectionofequations3.1and3.2,thefollowingconclusionscanbedrawn.
■ Thegreaterthepressuredropperunittime,thesmallerthemaximumdropletsizethatwillbegenerated.
Thusapressuredropexperiencedoverashorttime(e.g.overacontrolvalve)willgenerateasmallerdrop-
letsizethanthesamepressuredropexperiencedoveralongertime(e.g.pipeline).
■ Thelowertheinterfacialtension,thesmallerthemaximumdropletsizethatwillbegenerated.

TherelationshipdescribedbytheHinzeequationisillustratedgraphicallyinFigure3.1forarangeoftypi-
calvaluesforanoilproductionfacility.ThemixingenergiesinFigure3.1rangefrom0.0010m2/s3which
isthemixingenergyassociatedwiththefillingofsettlingtanksupto100,000m2/s3whichisthemixing
energythatmaybeimpartedoveracontrolvalve.Itcanbeseenthatthestabledropletsizeoverthisrange
fallsfromtheorderofmillimetersforthelowmixingenergytotheorderof10µmforhighmixingenergies.A
stabilizeddispersionwouldtendtogenerateevensmallermaximumdropletsizesduetothelower(effective)
surfacetensionandinhibitionofcoalescence.

Fig. 3.1. Relation between mixing intensity and maximum stable droplet size.

3.4. Droplet size distribution


Thedropletsizedistributionpertainingtoacertainmaximumdropletsizecanbeapproximatedbyastandard
distribution,whichcanbedescribedconvenientlybyanequationoftheform:

d
Vcum≈(1-exp(−3[]2)x95% (3.3)
 dmax
Thisistheso-calledRosinRammlerequation.ThisequationisshowninFig.3.2.foradispersionwithamaxi-
mumdropletsizeof100µm.Itcanbeseenfromthisfigurethattoremovethisdispersionwithanefficiency
of90%alldropletslargerthan20µmmustberemoved.Alternativelyitcanbecalculatedthatwhenthis
dispersionwouldbepassedthroughahydrocyclonewithadropletcutoffsizeof5microntheseparation
efficiencywouldbe>99%.


42 DeoilingManualV2.0
Restricted

100.0
90.0

Cumulative volume (%) 80.0

70.0

60.0

50.0
40.0

30.0

20.0

10.0

0.0
0 20 40 60 80 100.0 120.0

Droplet size (μm)

Fig. 3.2. Cumulative droplet size distribution pertaining to a d95 of 100 µm.

3.5.Effect of pressure drop over a valve on droplet size distribution and separation

Themaximumdropletsizeinadispersiondownstreamofavalveisafunctionofthepressuredropoverthe
valve, and the dispersed phase concentration. Extensive work on breakup in choke valves, recently pub-
lishedbyVanderZandec.s.attheUniversityofDelft,showsthatforlowconcentrationstheexperimental
datacanbedescribedsuccessfullywithacorrelationaccordingtoPercyandSleicher:  

       (3.3)

whered95isthedropletsizebelowwhich95%ofthevolumeisfound,doistheorificediameter,σtheinter-
facialtensionandDPmaxthemaximumpressuredropovertherestriction,whichisrelatedtothepermanent
pressure drop [2]. Equation (3.3) is basically a variation of the Hinze equation. Van der Zande found a
slightlylargerproportionalitycoefficient,i.e.C1=5.4insteadofthe3.1originallyreportedbySchleicher.
DPmaxinitsturnisalsorelatedtodo.

Fromthemaximumstabledropletsizethecorrespondingdropletsizedistributioncanbeestablishedbyas-
sumingastandardlognormalorRosinRammlerdistribution,see3.4.

Thecorrelationsintroducedinthischaptercanbeusedforexampletoestimatetheeffectofdepressurization
acrossavalveontheperformanceofdownstreamhydrocyclones.OneofthelinesinFig.3.3.showsthe
variationofthemaximumdropletsizedownstreamofthevalvewithpressuredropaccordingtoequation
(3.3).AswillbediscussedinChapter7thecutoffdropletsizeofawelldesignedhydrocyclonedependson
thedensitydifferencebetweenwaterandoil.Foralightfluidlikecondensate5micronmustbeachievable.

Thecondensateslipthroughthecyclonescanbeestimatedfromthevolumefractioninthefeedsmallerthan
thecutoffdropletsize.TheresultofthiscomparisonisalsoshowninFig.3.3.Itcanbeseenthatahigh
pressuredrophasahugedownsideintermsofwaterquality.

DeoilingManualV2.0 43
Restricted

Fig. 3.3. Effect of pressure drop across a valve on maximum droplet size and resulting condensate slip.

3.6. References

1. J.O . Hinze, ‘Turbulence’, Mc Graw-Hill, 2nd ed. (1975).


2. M.J. van der Zande and W.M. van den Broek, ‘Break-up of oil droplets in the production system’, Proc.
Energy Sources Technol. Conf., Houston, February 1998.

44 Deoiling Manual V 2.0


Restricted

Chapter 4.

Water disposal

Deoiling Manual V 2.0 45


Restricted

4. Water disposal
4.1. General
The water produced as a result of production operations needs to be disposed in a responsible and environ-
mentally acceptable manner. There are three major forms of water disposal.
■■ Surface disposal
■■ Water injection
■■ Disposal injection

Surface disposal is subject to regulation and is a potential source of environmental pollution. Regulatory re-
strictions, as well as a Group commitment to minimising discharges into the environment, will see subsurface
disposal techniques become more important in future operations.

The different water disposal methods set the performance required from deoiling equipment, such as the
regulatory limits on hydrocarbon concentration levels, (in particular) for surface water disposal, or reservoir
imposed limitations on solids and hydrocarbon content imposed for subsurface disposal.

4.2. Surface disposal – regulatory limits


Surface disposal of water involves the discharge of treated effluent water into the surface environment.
Because of the potential impact on the environment, the discharge of effluent water is normally governed by
regulatory limits set by national or regional authorities.

In general the regulations covering water discharges are most stringent for inland locations, with the regula-
tions often controlling a wide variety of substances including both dispersed and dissolved hydrocarbons,
heavy metals, biological oxygen demand etc.

In contrast, regulations covering offshore effluent discharges are less coherent. Limits have typically been
set on the basis of expected performance levels of available deoiling equipment and not on environmental
impact considerations.

It should be noted that regulations from different countries often have considerable variation in areas such
as defining which components are considered hydrocarbons, whether dissolved and polar hydrocarbons
are included and which hydrocarbon analysis method should be used. This is discussed in more detail in
Chapter 5.

The Shell HSSE & SP Control Framework (version 2, December 2009) specifies the following discharge limits:
■■ For existing assets the oil content of produced water (weighted monthly average of daily measurements)
should not exceed 30 mg/l for discharges in the open sea, and 15 mg/l for discharges in coastal zones.
■■ For new major installations for onshore oil and gas development the total hydrocarbon content of water
discharged should not exceed 10 mg/l .
■■ For new major installations for offshore oil and gas development the oil content of produced water (weight-
ed monthly average of daily measurements) should not exceed 30 mg/l, and daily measurements should
not exceed 42 mg/l..

The Group Standard requirements of 15, 30 and 42 mg/l are defined and measured by either:
■■ OSPAR Agreement 2005-15 ‘OSPAR Reference Method of Analysis for the Determination of the Dispersed
Oil Content in Produced Water’, which is required to be used for all OSPAR member countries;
or:
■■ United States Environmental Protection Agency (US EPA) Test method 1664 ‘Guidelines Establishing Test
Procedures for the Analysis of Oil and Grease and Non-Polar Material’.

Sampling and storage shall be in accordance with ISO 5667-3 ‘Water quality - Sampling - Part 3: Guidance
on the Preservation and Handling of Water Samples’ and applicable regulatory requirements.

46 Deoiling Manual V 2.0


Restricted

Apart from these limits EP Companies shall ensure that water discharges comply with local regulations includ-
ing restrictions on creating visible sheens. Residual concentrations of oil in water do not generally produce
visible sheens under normal operating conditions. However it may be necessary to apply chemicals in order
to eliminate sheens. This method shall only be applied if it can be demonstrated that it does not increase the
overall level of environmental impact and is in compliance with local regulations.

4.3. Water injection


4.3.1. General
For the purposes of this manual, the term water injection is used to mean the injection of waste water into the
production reservoir for the purpose of assisting hydrocarbon recovery. In some cases, only the production
water component of the waste water is considered suitable for water injection, with the remaining waste
water disposed using alternative methods.

The term disposal injection is used to define the injection of waste water into non-producing formations for
the purposes of disposal only. This is discussed separately in section 4.4

The injection of water into a production reservoir to assist the recovery of oil reserves is a relatively well
established procedure. Both surface water (sea, river and lake water) and subsurface water (aquifer and
produced water) have been satisfactorily injected in a variety of locations.

The following discussions briefly introduce the water treatment considerations that need to be addressed to
ensure the satisfactory injection of water as a means of both hydrocarbon recovery and waste water dis-
posal. The discussions focus on the injection of produced water which is normally the largest waste water
stream, however the general principles apply to most water sources.

For more detailed information on water injection the Waterflood Manual [1] takes precedence over this
manual.

4.3.2. Water compatibility


Water injection of produced water is only possible if the water is determined to be compatible with the
production reservoir. The investigation of water compatibility should address the following considerations:

■■ Compatibility with the formation. Chemical incompatibility may result in the expansion of clay minerals or
oxidation reactions due to the presence of oxygen (most injection water is maintained oxygen free). Physi-
cal incompatibility may result in the migration of fine solid particles through the formation (either carried by
the injection water or dislodged from the formation) with the potential for blocking pores and consequent
reduction in formation permeability.

■■ Compatibility with formation fluids. Incompatibility with formation fluids can result in changes in the chemi-
cal equilibrium of the formation fluids, potentially leading to scale formation in the reservoir, tubing or in
surface production facilities.

■■ Compatibility with other injection water. In some cases, waste water will form only part of the total volume
of required injection water. In this instance the compatibility of the waste water with the other injection
water (e.g. sea water) must be established, as well as the compatibility of the combined water with the
formation itself and the formation fluids.

Deoiling Manual V 2.0 47


Restricted

4.3.3. Permeability impairment


Treatment of water before injection is required to avoid potential impairment of the permeability of the forma-
tion. Table 4.1 lists potential sources of permeability impairment. Additional information can be found in the
Waterflood Manual.

The removal of suspended solids and dispersed hydrocarbons from the produced water are the most impor-
tant for the selection of water treatment equipment.

4.3.4. Solids removal


Solids present in injected water have the potential to be trapped in the formation around the well bore, result-
ing in a flow restriction and a deterioration in injectivity. Such blockage will adversely affect the ability of the
injection well to accept the requisite waste water volumes.

Individual reservoirs differ in tolerance for both the quantity and the size distribution of the solids in the injec-
tion water. Factors that will affect the ability of the reservoir to tolerate solids include;
■■ The permeability and pore structure of the basic reservoir material.
■■ The degree of fracturing present in the reservoir.
■■ The physical properties of the hydrocarbons present in the reservoir.

Core flooding tests can give some indication to the tolerance of the reservoir to the presence of solids. How-
ever field experience often contradicts the results of core flooding tests, either by injectivity impairment of
“tolerant” reservoirs or in the unexpected tolerance of supposedly “tight” reservoirs.

Typical quality requirements for water injection of sea water in the North Sea are 98% removal of solid
particles larger than 2 μm with a particle number count of less than 25 particles per 0.05 ml of water.
This corresponds to the typical water quality that can be achieved with conventional filtration equipment. A
well-head guard filter can be used to achieve a further reduction in solids by the removal of up to 100% of
particles larger than 0.6 μm.
However, the setting of excessively strict water injection specifications is very expensive in terms of additional
process equipment, particularly for offshore installations. Specifications should not be set as simply the best
achievable quality, but should be justified by consideration of all relevant information. Issues such as down
hole contamination of the water with corrosion/erosion products, the thermal fracturing of reservoirs and the
cost/benefit analysis of different treatment options should be considered.

4.3.5. Hydrocarbon removal


The presence of dispersed hydrocarbons in the injected water can also lead to permeability impairment.
Dispersion of the hydrocarbon into the reservoir may result in the formation of a stable emulsion or sludge,
often stabilised by the presence of solids. This emulsion may exhibit a very high viscosity and non-Newtonian
rheological behaviour, resulting in plugging of the reservoir.

In addition, dispersed oil may alter the wetting properties of the formation and may form an oil bank in the
proximity of the well bore, also reducing permeability.

For many water injection schemes the dispersed hydrocarbons are removed along with the suspended solids
by filtration. In this case the hydrocarbon droplet size in the feed stream and the level of dissolved hydrocar-
bons present will determine the level of hydrocarbons remaining in the filtered water. Typical design levels
are less than 5mg/l of dispersed hydrocarbons.

4.3.6. Secondary waste water streams


While water injection can be an efficient method for disposal of waste water, particularly produced water,
consideration must also be given to the disposal of secondary waste water streams from the production facili-
ties such as process water, drains water and ballast water.

48 Deoiling Manual V 2.0


Restricted

Table 4.1. Potential sources of permeability impairment

Impairment source Pre-engineering evaluation Method of prevention


method
Suspended solids ■ Source water survey ■ Fine filtration
■ Evaluation of filtration specification ■ Fine filtration
Coalescing oil droplets ■ Source water analysis ■ Fine filtration
■ Emulsification

Formation of insoluble scales ■ Analysis of source water and ■ Prevent commingling of

formation water incompatible fluids or use


■ Computer aided scale prediction chemical inhibitors
■ Dynamic tests of chemical inhibitor ■ Sulphate removal

membranes
Swelling of reservoir clays ■ Core sample flooding tests scales ■ Modify water injection

■ Petro-graphical analysis program


■ Chemical injection

Movement of reservoir fines ■ Core sample flooding tests ■ Modify water injection

■ Petro-graphical analysis program


■ Chemical injection

Bacteriological solids ■ Source water analysis ■ Inject chemical biocides

Corrosion products ■ Source water analysis ■ Materials and coatings


selection
■ Chemical inhibitor injection

■ Cathodic protection

In some cases, it may be possible to mix secondary waste water with the water for water injection. However
the suitability of this must be carefully evaluated as it may potentially introduce more problems than it solves,
due to;

■■ Thepossibility of scale formation from mixing incompatible water


■■ Theintroduction of bacteria and/or oxygen into the injection water.
■■ Mismatch of the optimum rate of water injection for production and the rate required for disposal of sec-
ondary waste water. During initial years of production there may be no requirement for any water injection.
■■ Seasonal variation of the flow rates of secondary water.

In most instances, water for injection and secondary sources of water are not mixed. In these circumstances,
even though the bulk of the waste water may be disposed of through water injection, waste water treatment
facilities must still be provided for the disposal of secondary waste water streams.

4.4. Disposal injection


4.4.1. General
Disposal injection differs from water injection in that the water stream is injected only for the purpose of
disposal, not as a means of assisting production rates. Disposal injection injects waste water into subsurface
formations other than the production reservoir, such as aquifers or exhausted production reservoirs.

Deoiling Manual V 2.0 49


Restricted

As with water injection, this Manual focuses on on the aspects of disposal injection relevant to the selec-
tion and design of deoiling equipment. For more detailed information the user is referred to the Waterflood
Manual.

4.4.2. Water quality constraints


Reservoir/formation constraints
The drilling and completion of any well is expensive and the water for disposal should be treated to a quality
sufficient to ensure that the capacity of the disposal formation is not restricted by avoidable damage. There-
fore, until proven otherwise, the constraints imposed by the reservoir on disposal water are broadly similar to
those outlined for water injection in section 4.3.

However, there is typically more latitude allowed in the treatment of disposal water than allowed for injection
water and there are numerous cases of the disposal injection of waste water with minimal treatment. A more
relaxed water quality specification for disposal water is often acceptable because:

■■ The disposal formation may be specifically chosen on the basis of a high tolerance for poor quality disposal
water.
■■ Stimulation of the disposal formation may be possible.

In many cases, water for disposal injection may only need to be treated to level similar to typical offshore
surface disposal specifications, i.e. below 30 mg/l dispersed hydrocarbons. This can be used as a starting
point for design.

However, considering the potential impact on equipment design, these limits should be confirmed through
normal reservoir plugging evaluations as discussed in the Waterflood Manual.

Environmental constraints
Disposal injection of waste water should avoid subsurface structures that may be used or linked to sources of
water for industrial, agricultural or domestic purposes.

Regulatory constraints
Countries which utilise aquifers as a source of water for industrial, agricultural or domestic purposes will often
have regulations governing the disposal of waste water in subsurface structures.

4.5. SECONDARY (REJECT) STREAM DISPOSAL


4.5.1. General
The main focus of most deoiling system designs is the quality of the effluent water stream. However, the
contaminants removed from a waste water stream are simply moved from the waste water stream to some
secondary waste stream. The design of a water treatment facility must consider the potential disposal require-
ments of secondary waste streams that may be generated.

Secondary waste streams may include concentrated hydrocarbon streams, sludge and waste gas streams.
In particular, the techniques used for the removal of dissolved hydrocarbons often result in the generation of
large secondary streams that must be treated, regenerated or disposed.

4.5.2. Disposal options


Export with hydrocarbons
Secondary waste streams may often be spiked into the product hydrocarbon stream and exported. This is
often possible as the secondary waste stream may be relatively concentrated and small in volume compared
to the overall product stream. This is a common method of “disposing” the hydrocarbons recovered from
deoiling processes.

50 Deoiling Manual V 2.0


Restricted

However, it should be recognised that this method of disposal is simply transferring potential contaminants
to downstream processing activities where the same contaminants may need to be removed a second time.
In some cases it may be more economic to treat the waste upstream rather than transfer the problem down-
stream.

Recycling to process
Many process schemes recycle secondary waste streams to the start of the process. This is satisfactory pro-
vided at some point in the process there is a satisfactory bleed of the contaminant from the process.
Care must be taken to avoid the potential accumulation of contaminants.

The recycling of sludge, such as rag layers from oil/water interfaces, should be avoided. Such sludges are
relatively stable and are unlikely to break down without additional treatment. As a result they will tend to ac-
cumulate in the process. Solids may also accumulate in the process and may eventually reach a stage where
they stabilize sludges or emulsions.

Disposal to atmosphere
Some deoiling processes will result in the generation of a waste gas stream, for example gas flotation. The
resulting gas stream containing stripped hydrocarbons is typically discharged to the flare system or fuel gas
system, and may be subject to regulations governing discharges to atmosphere.

Disposal for separate treatment


In some circumstances, secondary waste streams will be removed from the process for separate treatment.
The two best examples are emulsion sludges and solid sludges.

Emulsion sludges are stable oil/water emulsions, often accumulating at the oil/water interface. Such emul-
sions are stabilised by untreated indigenous surfactants in the crude, production chemicals, waxes or fine
solids. These emulsions are to be withdrawn for separate treatment in dedicated sludge treatment plants.

Solids sludges are secondary waste streams containing accumulated solids, for example solids removed by
wellhead desanding systems or washed out of separators. In onshore operations this may include organic
material and biomass .
The problematics of sludge streams should not be underestimated. They may contain concentrated levels of
environmentally significant components such as hydrocarbons, heavy metals and radionuclides and may be
subject to special environmental regulations. Adequate provision must be made for the removal, handling,
storage and treatment of solid and sludge waste products at the design stage.

The Sand Management Manual EP 2003-5184 deals extensively with the handling and disposal of sand
and solids sludges.

4.6. References
1. http://sww.shell.com/ep/technology/ped/waterflooding/waterflood_manual.html

Deoiling Manual V 2.0 51


Restricted

52 Deoiling Manual V 2.0


Restricted

Chapter 5.

Analysis and monitoring of


waste water

Deoiling Manual V 2.0 53


Restricted

5. Analysis and monitoring of waste water

5.1. General
For most E&P deoiling operations, two of the most important characteristics of the water stream are the
concentration of the dispersed and dissolved hydrocarbons and the size distribution of the dispersed hydro-
carbon phase.

Infrared absorption analysis is the most common procedure for determining the concentration of hydrocarbons
in waste water and the advantages and limitations of this method of analysis are presented in this chapter.
Of particular importance is the recognition of the following potential limitations in many infrared analysis
procedures;
■■ The absorption frequencies used may not be sensitive to the presence of aromatic hydrocarbons.
■■ The standard procedures do not differentiate between dispersed and dissolved hydrocarbons.
■■ The accuracy of the measurement is dependent on the use of suitable calibration samples.

Infrared adsorption analysis is discussed in detail in section 5.3. However other methods of analysis are
also available and these are briefly discussed in section 5.4. On-line oil-in-water monitors are discussed in
section 5.5.

The size distribution of the dispersed hydrocarbon phase is one of the most important parameters governing
the performance of most deoiling equipment. The measurement of droplet size distributions is discussed in
section 5.6.

5.2. Oil in water


it is not straightforward what is meant by oil in water. Depending on the specified regulatory test method, Oil
in Water (OiW) may be composed of only a fraction of the dispersed or dissolved hydrocarbons. There may
be other constituents as well such as organic acids, or even injected chemicals. The oil components that tend
to dissolve in water are the polar components such as BETX (benzene, ethylbenzene, toluene, and xylene),
NPD (naphthalene, phenanthrene, thiophene, dibenzothiophene), polynuclear aromatics, phenols, resins,
short chain acids, and naphthenic acids. Oil components can be dispersed in the form of drops in the pro-
duced water. Oil drops might accumulate surface active components such as acids, asphaltenes, and resins.

The amount of dissolved and dispersed organics in the produced water can increase or decrease depending
on the processing conditions, such as temperature, flow rate, and on which wells are producing at any given
time. Some of these components may be present but might not contribute to the measured OiW. The relative
contribution that these components make to the OiW content depends on the method used for testing.

A distinction should be made between regulatory tests and on-site monitoring tests. The testing required to
meet the local regulations is often different from the monitoring methods that are used. Obviously they have
to be related in some way such as through a correlation. Typically, the regulatory tests are carried out at
a qualified lab and are more involved, take more time, and require more specialized equipment than the
monitoring methods. In the UK/European sector, regulatory tests are carried out using gas chromatography
on some of the larger production facilities. Whereas an operator has little leeway to vary the regulatory test
methods, operators generally have a wide range of options for monitoring methods. In any case, the operator
must understand the relation between the regulatory tests and the production monitoring methods.

Monitoring methods vary widely from one region to another and from one asset to another. One reason is
that regulatory test methods differ significantly from one region of the world to another. The second is the
availability of local vendors to supply, repair, and help calibrate the monitoring equipment used in the field.
The best monitoring method in one part of the world may be completely unsuitable in another.

54 Deoiling Manual V 2.0


Restricted

5.3. Oil in water measurement – infra red absorption methods


the most common method of measuring the concentration of hydrocarbons in water is infrared absorption.
This method is based on the absorption of infrared light by the carbon-hydrogen (C-H) bonds present in hy-
drocarbons. The procedure generally consists of the following steps.
■■ Sample preservation and preparation.
■■ Extraction of the hydrocarbons from the water using an organic solvent.
■■ Conditioning of the extract to remove solids, water and polar hydrocarbons.
■■ Measurement of the infrared absorption at one or a number of infrared light wavelengths.

A detailed description is given in Appendix 5.1.

Standard infrared analysis procedures measure the total quantity of hydrocarbons extracted from the water
by the solvent and make no distinction between dispersed and dissolved hydrocarbons.

The results obtained from infrared analysis are not an absolute measure of the hydrocarbon content of effluent
water, but are calibrated against reference standards. Inappropriate calibration standards will result in an
inaccurate determination of hydrocarbon content. In particular, the aromatic content of effluent samples may
be excluded from the analysis due to inappropriate calibration standards or the use of infrared wavelengths
which are not absorbed by the carbon-hydrogen bonds present in aromatics.
In addition, many infrared analysis procedures do not measure polar hydrocarbons which are removed from
the sample before analysis. This includes both polar hydrocarbons derived from the well-stream (organic
acids and phenols) and polar process chemicals such as methanol and glycols.

It should be noted that infrared hydrocarbon analysis is only a relative determination of hydrocarbon
content in water. When a more complete breakdown of the hydrocarbons in the effluent water is
required a more absolute method of analysis such as gas chromatography must be used.

5.4. Oil in water measurement - other methods


5.4.1. Dispersed and dissolved hydrocarbons
Most deoiling equipment installed at E&P sites is only capable of removing dispersed hydrocarbons. Being
able to differentiate between dispersed and dissolved hydrocarbons may be necessary to evaluate the per-
formance of deoiling equipment.

The most common methods used for the determination of hydrocarbon levels in waste waters involve the
extraction of the hydrocarbons from the water phase using a solvent. Both dispersed and dissolved hydro-
carbons will be extracted by the solvent and the total hydrocarbon content of the sample will be determined,
not just the dispersed hydrocarbons.

The simplest method to differentiate between dissolved and dispersed hydrocarbons is to perform the
infrared analysis procedure on two effluent samples. The first effluent sample is analysed normally, thus
measuring the total hydrocarbon content.

The second effluent sample is filtered through a suitably fine filter to remove the dispersed hydrocarbons. The
filtrate will then only contain the dissolved hydrocarbons which can be measured using the normal analysis
procedure. The dispersed hydrocarbon content can then be calculated by subtracting the dissolved hydrocar-
bon content from the total hydrocarbon content.

5.4.2. Gravimetric analysis


A solvent is used to extract the hydrocarbon content from the effluent sample. The extract is filtered to remove
solids, water and salt, then evaporated under standard conditions to leave a residue. The residue is weighed
and is reported as the hydrocarbon content.

Deoiling Manual V 2.0 55


Restricted

Gravimetric analysis tends to underestimate the concentration of hydrocarbons in water due to the loss of
lighter hydrocarbons during the evaporation procedure. Comparative trials indicate that the gravimetric
method may measure a hydrocarbon concentration in the order of 20% lower than infrared analysis.

However, gravimetric analysis has the advantage that, apart from evaporation losses, it is an absolute analy-
sis and is independent of calibration errors.

5.4.3. Visible spectrum colorimetry


A solvent is used to extract the hydrocarbon content from the effluent sample. The extract is then measured for
absorption in the visible light spectrum or is visually compared with fresh reference samples of hydrocarbons
in solvent.

This method has the advantage that dissolved gases and many other non-oily organic compounds, being
colourless, do not interfere with the measurement. Other advantages are;
■■ Solvents containing carbon-hydrogen bonds can be used.
■■ It is not normally necessary to filter, dry or pass the sample through an absorption column.
■■ The method is simple, cheap and portable and suitable for use under difficult conditions.

However, the method may have a tendency to underestimate hydrocarbon content as the coloured compo-
nents in crude oil are very easily absorbed onto surfaces and may be missed.

5.4.4. Gas chromatography


The hydrocarbon content of the effluent sample is extracted with a suitable solvent and injected into a gas
chromatography column which separates hydrocarbons in order of boiling point. Gas chromatography gives
a detailed analysis of the composition and may be applied for special studies, but is generally too time
consuming for routine use.

The main disadvantage of this method is that high boiling point components (>300 oC)are not detected as
they are not carried through the column. The sensitivity of the method is 0.1 mg/l for oil, for single compo-
nents less than 0.01 mg/l.

5.4.5. Others
Ultraviolet absorption
Aromatics and other compounds with conjugated or pseudo-conjugated double bonds will absorb UV light.
If the typical concentration of these particular hydrocarbons is known, the UV absorption can be related to
the overall hydrocarbon content. Dual wavelength UV beams are sometimes used, the first measuring the
absorption of the aromatics and the second acting as a reference to compensate for fouling and turbidity.

Ultraviolet fluorescence spectrometry


Aromatics and other compounds with conjugated or pseudo-conjugated double bonds will also fluoresce
when exposed to suitable ultraviolet light. The fluorescence can be related to the hydrocarbon content of the
effluent sample.

Concentration derived from droplet size analysis


Dispersed oil-in-water takes the form of discrete oil droplets suspended in the continuous water phase. If
the number and size of the hydrocarbon droplets is known, the volume of hydrocarbon in the water sample
can be determined by calculating the cumulative volume of the hydrocarbon droplets. Knowing the average
density of the dispersed hydrocarbons present allows the total hydrocarbon concentration to be calculated.

Ultrasonic oil-in-water monitors operate on the basis of recognising the presence of dispersed droplets. An
ultrasonic signal is reflected back by dispersed hydrocarbons in the sample to give an indication of the hy-

56 Deoiling Manual V 2.0


Restricted

drocarbon content. The result will be affected by the droplet size, therefore the sample must be homogenised
before analysis.

Droplet size analysers are available which allow continuous on-line measurement of droplet size distributions.
This method will only measure the concentration of dispersed hydrocarbons, dissolved hydrocarbons will not
be measured. There is also potential for interference from gas bubbles and solid particles and the technique
will become more inaccurate if a significant quantity of the hydrocarbons are present as very small droplets
(< 5 mm).

5.5. On-line oil-in-water analysers


the development of an on-line oil-in-water monitor has been pursued by a number of manufacturers. Con-
tinuous measurement of the hydrocarbon content in the effluent waters from production sites would have the
following potential benefits.
■■ More rapid detection of excess hydrocarbon content in effluent discharges.
■■ More feedback on the operation of deoiling equipment, promoting improvements in operational proce-
dures and equipment design.
■■ Minimisation of manual sampling and analysis requirements.

A range of commercial oil-in-water monitors have been tested by KSEPL and in trials at the Orkney Water Test
Centre (OWTC). Although some of the meters tested have shown potential, at the time of testing none could
be recommended as suitable for continuous operation in an operating environment.

However, it should be recognised that the suitability of continuous oil-in-water monitors is often gauged on
their absolute accuracy and their potential to replace current sampling and analysis procedures. Although
commercial monitors may not yet be suitable for this purpose, their use in less rigorous applications should
not be overlooked. Commercially available monitors could potentially be used in the following applications:
■■ To give a continuous indication of relative deoiling performance, calibrated by continued routine sampling
and analysis.
■■ To act as an alarm, indicating when significant excursions occur in effluent hydrocarbon levels. This would
facilitate more rapid corrective action, minimising the potential impact on downstream processes or the
discharge environment.

5.6. Measurement of droplet sizes


5.6.1. General
One of the most important criteria governing the selection and performance of deoiling equipment is the
droplet size distribution of the dispersed hydrocarbons in the feed stream. As a result, the ability to accurately
measure the droplet size distribution is a vital tool in the correct design and operation of deoiling equipment.

This is particularly important when upgrading existing equipment, where the existing droplet size distribution
can be measured and the optimum deoiling equipment can be selected.

5.6.2. Sampling
The droplet size distribution of a dispersed hydrocarbon in water is constantly changing as a result of the
dispersing and coalescing forces acting on the fluid. To accurately measure the droplet size distribution a
careful sampling procedure must be followed to ensure that the sample is representative. Obtaining a repre-
sentative sample of a two phase liquid can be very difficult and a number of samples may be required with
the results averaged.
■■ The sample point should be correctly designed with a quill extension into the main fluid flow away from the
piping wall. In some cases samples may be required at different points to ensure that the droplet distribution
does not alter over the cross-section of the area being sampled (e.g. stratified flow).

Deoiling Manual V 2.0 57


Restricted

■■ Iso-kineticsampling should be used at the point where the sample enters the sample system. This ensures
the sample is not biased by changes in the flow profile.
■■ Iso-energetic conditions are to be maintained after the sample has entered the sample system. This will en-
sure that sampling does not alter the droplet size distribution by changing the dynamic equilibrium between
coalescence and emulsification.

5.6.3. Potential problems in measuring droplet size distributions


In addition to the sampling considerations given above, the following points should also be considered when
measuring droplet size distributions.
■■ Most equipment for the measurement of droplet size distribution operates at atmospheric pressure, making
it impossible to avoid depressurisation of the sample. Depressurisation of the sample may have a significant
effect on the droplet size distribution.
■■ Some equipment for the measurement of droplet size distribution may be temperature sensitive, requiring
the sample to be cooled before measurement. This cooling may also affect the droplet size distribution.
■■ Coalescence may rapidly alter the size distribution of dispersed droplets. Samples must be analysed as
quickly as possible.
■■ Solid particles and gas bubbles present in the sample may be detected as droplets, leading to a skewed
size distribution.

5.6.4. Droplet size distribution analysers


A number of different methods are currently available for the measurement of droplet size distribution.

Aperture counting (Electrical zone sensing)


Aperture counting based droplet size analysers were developed by Coulter Inc. The dispersed hydrocarbon
droplets pass though a small, electrically charged aperture. The passage of the droplet results in a distur-
bance in the electrical field. From the number and magnitude of the electrical disturbances the size distribution
of the droplets can be determined.

The correct size of the aperture is linked to the size distribution being measured. If the aperture is too small,
deformation of the droplet may occur as it passes through the aperture. In some cases the sample must be
diluted with an isotonic liquid to achieve a suitable droplet population density, however this dilution may
influence the droplet size distribution.

Diffraction based methods


Diffraction based methods infer the droplet size distribution from the diffraction patterns caused by the interfer-
ence between suspended hydrocarbon droplets and a laser beam. Such instruments can typically measure
particle sizes down to the order of 0.1 µm.

Diffraction based methods are sensitive to the number of droplets in the sample. Too few droplets may not
give a statistically significant result while too many droplets may result in optical interference with the measure-
ment. Not all equipment warns the user that these errors may be occurring.

Time of transition
Time of transition based methods measure the interaction pulses generated as droplets intersect a scanning
laser beam. Analysis of the interaction pulses determines the droplet size distribution.

The interaction pulses are affected by the shape of the particle allowing differentiation between particles of
different shape. Thus solid particles such as sand or corrosion products, which tend to have asymmetrical
shapes can be differentiated from symmetrical hydrocarbon droplets.

58 Deoiling Manual V 2.0


Restricted

As with diffraction based methods, the number of droplets in the sample may influence the result. Too few
droplets may not give a statistically significant result while too many droplets may result in optical interference
with the measurement. Not all equipment warns the user that these errors may be occurring.

Deoiling Manual V 2.0 59


Restricted

60 Deoiling Manual V 2.0


Restricted

Appendix 5.1

Infrared absorption analysis

Deoiling Manual V 2.0 61


Restricted

Appendix 5.1. Infrared absorption analysis

A5.1.1 Sample preparation

Acidification
If the sample is not to be analysed immediately, it is common practice to acidify the sample. Acidification
has the following beneficial effects;
■■ Minimises hydrocarbon losses due to biological degradation.
■■ Improves phase separation by countering the presence of alkaline based emulsion stabilisers.
■■ Dissolves any precipitated calcium carbonate.

Hydrochloric acid is typically used to avoid potential precipitation problems that may occur with anions from
other acids (e.g. sulphates from sulphuric acid). ASTM D 3694-92, “Standard Practice for Preparation of
Sample Containers and for Preservation of Organic Constituents“, recommends acidification to a pH of 2 to
inhibit biological activity for biodegradable organic chemicals.

Acidification has a potential problem in that it may change the dissociation equilibrium of weak organic
acids. These liberated organic acids would then be detected by the subsequent hydrocarbon analysis and
incorrectly reported as part of the hydrocarbon component of the water.

This problem can be avoided by ensuring that after the organics are extracted from the water phase, the
extract is passed through a suitable polar absorbent. As the liberated organic acids are polar, they will be
removed by the absorbent and will not be measured as part of the mineral hydrocarbon content of the water.

Gas purging
The Paris Commission (Parcom) guidelines governing water discharges from offshore platforms in the North
Sea allows samples to be purged with a nitrogen gas flow (at approximately 1 litre/min) for a period of 45
seconds at a temperature of 30 °C.

The justification for this gas purging is to remove light components that would normally flash into the gas
phase on discharge into the environment, leaving only the hydrocarbons that will remain in the aqueous
environment. However gas purging will also strip volatile hydrocarbons from the sample. Up to 75% of the
BTEX aromatics may be stripped by this purge, resulting in an underestimate of the hydrocarbon content of
the effluent stream.

A5.1.2. Hydrocarbon extraction

Selection of solvent
The solvent used to extract the oil from a water sample must be invisible in infrared. This means that it must not
have C-H bonds (or other absorbance in the range of the aliphatic CH-H stretch of 2.93 µm). For this reason,
pentane and n-hexane of course cannot be used. As a consequence, any IR monitoring used in the Gulf of
Mexico and in the North Sea, where EPA 1664 and OSPAR apply respectively, will have to be implemented
with proof of correlation between the monitoring method and the regulatory test.

The obvious candidate solvents for IR monitoring are the Freon type solvents (such as Freon-113) which un-
fortunately are ozone depleting substances and have therefore been banned by the UN Montreal Protocol.
Consequently a solvent that was not on the UN shortlist for ozone depleting substances was required. Tet-
rachloroethylene was chosen through a process of elimination as the solvent of choice in the UK in around
1998. But the safe handling of the extraction solvent is an area that must be managed. Tetra-chloroethylene
is a potential carcinogen. Typical control measures include gloves, fume cupboard. Handling of the spent

62 Deoiling Manual V 2.0


Restricted

solvents must be considered. Usually, the spent solvents are processed onshore by a waste management
company so no discharge of the solvents occur either.

Extraction
Extraction of the hydrocarbons from the waste water is achieved by adding the solvent to the waste water
sample and ensuring intimate mixing of the phases. There is quite a wide variation in extraction procedures
between the methods presented in Appendix C. Key differences are:
■■ The use of mechanical mixing or manual mixing. Mechanical mixing is preferred to ensure a repeatable
result.
■■ The German DIN, Dutch NEN and Swedish standards utilise a centrifuging step after mixing to ensure
complete separation of the water and organic phases.

The extraction procedure does not differentiate between dissolved and dispersed hydrocarbons present in
the water sample. Both will be extracted from the water sample and measured in the subsequent infrared
analysis.

As dispersed hydrocarbons in the sample will tend to coalesce over time, it is difficult to accurately subdivide
a waste water sample. Thus the extraction must be performed on the entire water sample. The extraction may
be performed either in the sample container itself or else the sample container and the sample container lid
or stopper should be thoroughly rinsed with the solvent to ensure all residual hydrocarbons are extracted.

Filtration and drying


To remove solids and water droplets that may be present it is normal practice to filter the solvent extract. All
final traces of water may be removed by passing the extract though an anhydrous drying column containing
sodium sulphate or other suitable water adsorbent.

Removal of polar compounds


Polar hydrocarbons, including compounds such as organic acids and phenols, are normally present in the
waste water streams from oil and gas production. Their polar nature makes them relatively soluble in water
and they may have a relatively high concentration in effluent water streams.

The solvent extraction procedure extracts both mineral hydrocarbons and a proportion of the polar hydrocar-
bons from the water sample. Regulations governing hydrocarbon levels in effluent streams generally exclude
(or are interpreted to exclude) polar and dissolved hydrocarbons. Thus the polar hydrocarbons are often
removed from the solvent extract before measuring the hydrocarbon content.

Polar hydrocarbons are removed by passing the extract though a column containing a polar absorbent such
as Florisil (activated silica-gel) or alumina (A1203). However, it should be noted that this absorbent may also
remove some mineral hydrocarbons with polar characteristics from the sample, such as high molecular weight
aromatic compounds.

A5.1.3 Infrared analysis

Basis of measurement
Infrared analysis is based on the absorption of infrared light by the C-H bonds present in hydrocarbons, with
the absorption being related to the concentration of hydrocarbons.

The solvent extract obtained from the previous procedures is scanned over a range of infrared light wave-
lengths in an infrared spectrophotometer. The absorption at characteristic wavelengths is related to the pres-
ence of particular C-H bonds.

Deoiling Manual V 2.0 63


Restricted

The presence of any of these groups will result in some absorption, hence the need to remove polar hydro-
carbons before the absorption is measured. Similarly, the solvent used for the hydrocarbon extraction must be
selected on the basis of not contributing to the infrared absorption, hence the use of halogenated solvents.

Many infrared analysis procedures only measure absorption for the CH2 group. This is generally satisfactory
as this is the most common carbon-hydrogen group present in mineral oils.

However, aromatic hydrocarbons consist largely of CH bonds. If the absorption is not measured at the wave-
length absorbed by these bonds (3.3 µm), these compounds will not contribute to the absorption and may go
undetected (though this also depends on other factors such as the composition of the calibration sample). This
could lead to a underestimate of the hydrocarbon content in effluent waters, particularly when measuring the
hydrocarbon discharges from gas production operations where the dissolved aromatic content of the effluent
may be quite high.

Calibration
Once the absorption of the extract has been measured it must be converted into the required analytical
measurement. This is achieved by comparison with the absorption measured for a known oil or reference oil
calibration sample.

Known oils
Known oils are calibration samples made by mixing water with known quantities of a hydrocarbon sample
taken from another part of the process. A typical source for the known oil may be the hydrocarbon reject
stream of a water treatment package.

Such known oils may not be completely representative of the hydrocarbons present in the effluent sample,
especially if the water treatment system has additional separation stages after the point where the known
hydrocarbon is obtained. The known oil may also be subject to a degree of “weathering” which may alter
its absorption.

However, in general, known oils are likely to be more representative of the actual hydrocarbons in the sample
than the alternative calibration methods.

Reference oils
Reference oils are calibration samples made by mixing known quantities of a defined mixture of hydrocar-
bons. As an example, the reference oil defined by the Netherlands NEN 6675 standard consists of 37.5%
iso-octane, 37.5% cetane and 25.0% benzene.

The inherent drawback in the use of a reference oil is that if the composition of the hydrocarbons in the efflu-
ent stream is significantly different to that of the reference oil the accuracy of the hydrocarbon measurement
will be reduced.

The absolute accuracy of the method will also be reduced if the hydrocarbons in the sample contain a sig-
nificant fraction of fully substituted carbon atoms i.e. carbon atoms with no C-H bonds. These fully substituted
carbons atoms will not contribute to the absorption and thus will not be accounted for unless the reference oil
used also contains a similar proportion of fully substituted carbon atoms.

The advantage of reference oils is that they may be specified as part of standard method, allowing the
same reference mixture to be used consistently in different locations. This at least provides a consistent basis
for comparison of results from different locations and allows for the checking of procedures by independent
authorities. However the accuracy of the result will be lower due to the variation between the actual hydro-
carbons in each sample and the hydrocarbons in the reference oil.

64 Deoiling Manual V 2.0


Restricted

Specific absorption coefficients


An alternative to using calibration samples is the use of specific absorption coefficients. These coefficients can
be built into equations which relate the measured absorption directly to the hydrocarbon content.
As with use of a reference oil, the disadvantage of this process is that it assumes the composition of the efflu-
ent stream is similar to that of the sample used to calculate the specific absorption coefficients. Again, carbon
atoms that are fully substituted with non-hydrogen bonds will not be measured and will be excluded from the
measured hydrocarbon content, This can be overcome by the use of correction factors, however such factors
must be empirically determined for each application.

Results
Units of measurement
The concentration of hydrocarbons present in water should generally be expressed using the unit of mg/l.
The basis for the ppm unit (parts per million) is often poorly (or not) defined and could be interpreted as either
a volume or mass basis.

In some cases, legislation may require the reporting of results in alternate units such as mg/kg, ppm(w) or
ppm(v).

Accuracy
It can be seen from the discussions above that the accuracy of measuring the concentration of hydrocarbons
in water will be dependent on a number of factors:
■■ Obtaining a representative sample.
■■ The effectiveness of the extraction stage. This is in turn governed by factors such as the solubility of the hy-
drocarbons in the chosen solvent, the method and period of agitation used during extraction, then number
of extraction stages and the effective separation of the solvent from the water phase.
■■ The potential loss of hydrocarbons during the removal of polar hydrocarbons from the extract.
■■ The measurement of the absorption at suitable wavelengths.
■■ The suitability of the calibration method.

The accuracy of the method is generally reported to be in the order of ±10%.

Deoiling Manual V 2.0 65


Restricted

66 Deoiling Manual V 2.0


Restricted

Chapter 6.

Equipment selection and


system integration

Deoiling Manual V 2.0 67


Restricted

6. Equipment selection and system integration


6.1. Introduction
in many installations, the majority of design or operating effort is concentrated on the hydrocarbon producing
systems with only relatively brief consideration given to the waste water treatment system. In particular, water
treatment facilities are often regarded as “end of pipe” solutions to a water treatment “problem”.

A preferable situation is where water treatment is considered as an integral part of the total process design.
Integration of water treatment equipment into the overall process system will result in a more efficient and cost
effective facility.

Dehydration and deoiling processes should be considered as a single system with the emphasis on comple-
mentary design. Examples where such integration has brought improvements in operation include;
■■ Minimisation of shear and promotion of coalescence in flow lines, pipe lines and piping systems upstream
of dehydration has improved the quality of both oil and water outlet streams.
■■ The use of plate packs in primary separators to improve the quality of both oil and water outlet streams.
■■ A gas/condensate installation with hydrocyclones operating at a high separator pressure (circa 120 barg)
to avoid depressurization and accompanying droplet shearing.
■■ A gas/condensate installation designed to avoid the use of corrosion inhibitors (which stabilise small
hydrocarbon droplets) and with the gas/liquid separator designed for effective three phase separation at
high pressure.
■■ The use of continuous dehydration designs in preference to batch dehydration has resulted in more consist-
ent and stable feed streams to downstream water treatment equipment and more efficient water deoiling.

These examples highlight the potential benefits that may be achieved by adopting a system approach to the
design of the overall production facilities, particularly the dehydration and deoiling processes.

This philosophy is examined in more detail in the following discussions. Section 6.2 presents guidelines to
assist the user in effective process design and equipment selection for water treatment systems, while section
6.3 discusses overall system optimisation and integration considerations.

6.2. Equipment selection and system design


Figure 6.1 on the following page presents a simple flow diagram illustrating the typical steps that should be
followed when selecting a water treatment system. Each of the steps is addressed in more detail in the discus-
sions in sections 6.2.1 to 6.2.9.

6.2.1. Sources and magnitude of water streams (1)


The source of the waste water will strongly influence its characteristics. For example, water streams from oil
production, gas production or ballast water systems would often require different water treatment facilities
due to differing water characteristics.

Clearly identifying the sources and magnitudes of the waste water streams will assist in the development of
more efficient water treatment facilities. For example, small, difficult to treat streams may be handled sepa-
rately rather than mixing the problem into the entire water stream.

68 Deoiling Manual V 2.0


Restricted

Figure 6.1 Flow diagram for water treatment system design

6.2.2. Identify contaminants in the waste water stream (2)


The full range of contaminants that may be present in the waste water stream should be identified. In addi-
tion to dispersed hydrocarbons, particular emphasis should be given to the presence of suspended solids,
dissolved hydrocarbons and treatment chemicals.

The majority of deoiling equipment used in E&P operations will only remove dispersed hydrocarbons. The
presence of dissolved hydrocarbons may require more sophisticated equipment to meet the quality require-
ments in the treated water.

The distinction between dispersed and dissolved hydrocarbons is particularly important for gas operations
where dissolved hydrocarbons may constitute a significant proportion of the hydrocarbon content in the waste
water.

A range of potential contaminants of waste water is discussed in section 2.3.

6.2.3. Identify treated water quality requirements (3)


A suitable disposal route for the treated water should be identified. Both surface and subsurface disposal
options should be considered.

Deoiling Manual V 2.0 69


Restricted

Health, Safety and Environmental considerations should take first priority in the setting of treated water
quality requirements. The potential environmental impact of water discharges should be evaluated for both
surface and subsurface disposal methods.

Once HSE requirements have been satisfied, technical and regulatory requirements should be considered.
Probable changes in environmental legislation over the design life of the installation should be identified and
their potential impact on the design examined.

The final target water quality will be the most stringent of the HSE, technical and legislative requirements. The
setting of water quality targets is discussed in more detail in chapter 4.

6.2.4. Select a suitable process location for the water treatment system (4)
Most water treatment equipment is located at the end of the process, often operating at or close to atmos-
pheric pressure. However, whenever possible, consideration should be given to locating deoiling equipment
further upstream in the process. Many types of deoiling equipment may be operated at elevated pressures
(e.g. hydrocyclones, plate pack interceptors, flotation, filters etc.)

This minimises the pressure drop between the water source and the water treatment equipment, thus mini-
mising shear forces and maximising hydrocarbon droplet sizes. If possible, shear inducing devices such as
control valves should be located downstream of water treatment facilities.

Waste water streams are typically combined from a number of sources, however in many cases the flow from
just one source may provide the bulk of the water. Consideration should be given to separate treatment at the
source of the bulk waste water stream.

An excellent example is a conventional three stage oil separation train. The bulk of the water is often sepa-
rated in the first stage separator at a relatively high pressure, yet all this water is usually let down to a low
pressure in order that the relatively small water streams from subsequent separators can report to a common
low pressure water treatment system.

A more effective installation may be to operate water treatment system immediately downstream of the first
separator at an elevated pressure. The bulk of the water would then be effectively treated without being
sheared through level control valves. The relatively low flows of water from other sources could be pumped
to the water treatment system using low shear pumps or treated in separate facilities.

Similarly, it may be found that the majority of contamination may come from just one source. Treating the
contaminated stream separately may be more efficient than mixing the streams then having to remove the
same contaminants from a larger stream.

Consideration should also be given to the process conditions that will affect the separation mechanism. The
separation of two phases following Stokes Law will be sensitive to a number of variables which are affected
by process conditions. In particular, the density differential and viscosity will be functions of the process tem-
perature and the selection of the process location should consider the potential effect on the phase separation
of the operating temperature.

6.2.5. Identify upstream methods of improving ease of water treatment (5)


The performance of the water treatment system will be largely dependent on the nature of the feed to the
system.

The processes upstream of the water treatment system should be reviewed to determine whether there is
scope to improve the water quality feeding the water treatment system. Improvements in water quality could

70 Deoiling Manual V 2.0


Restricted

encompass increases in dispersed droplet size, a reduction in hydrocarbon concentration, elimination of gas
or suspended solids, or the use of corrosion resistant materials to avoid the use of corrosion inhibitors.

Improvements could be implemented by means such as minimising pressure drop through the system, fitting
plate packs or coalescing devices to upstream vessels and improving process control. An example of im-
proved process control is automatic ratio control of a chemical injection system with production rate.

The optimisation of upstream processes is also addressed in section 6.3, System Optimisation and Integra-
tion.

6.2.6. Select the number of treatment stages (6)


The overall oil removal efficiency required is > 95% and there is no deoiling equipment that is capable to
operate at such a high efficiency, certainly not under the variation of flow and concentration occurring dur-
ing normal operation. Therefore the deoiling process normally requires two or three consecutive stages. This
allows the upstream stages to optimise the feed conditions for the downstream stages, removing surges in
flow rates or concentrations, and reducing the hydrocarbon concentration to below the maximum that can be
handled by the downstream process.

Water treating equipment can be broadly categorized into the four stages illustrated in Table 6.1
Table 6.1 Typical water treatment stages and associated equipment

6.2.7. Select the suitable deoiling equipment (7)


The most important variable influencing the selection of most water treatment equipment for the removal of dis-
persed hydrocarbon droplets is the hydrocarbon droplet size. The hydrocarbon removal efficiency of gravity
based separation equipment is largely dependent on droplet size, most equipment having a minimum droplet
size that can be efficiently separated.

Deoiling Manual V 2.0 71


Restricted

Unfortunately, at the design stage for a new installation, it is very difficult to predict the likely droplet size
that may be present in the feed to the water treatment equipment. Comparisons may also be made to the
performance of water treatment equipment or the measured droplet size distributions at existing installations
with similar process and operating conditions.

The design of water treatment equipment to be added to existing installations should be based on field
measurements of the hydrocarbon droplet size distribution. Such measurements give added confidence to a
particular design and should allow the selection of the most cost effective treatment scheme.

A number of water treatment equipment designs can be used to separate a given droplet size distribution.
Once the droplet size criteria has been satisfied the selection of the water treatment system is governed by
factors such as;
■■ Size and weight (particularly offshore)
■■ Capital cost
■■ Operating cost
■■ Ease of operation
■■ Maintenance requirements
■■ Flexibility of operation
■■ Secondary waste streams

Table 6.2 presents some guidance for the selection of different water treatment equipment designs. However,
for effective equipment selection all the steps presented in Figure 6.1, particularly water and droplet size
characterisation, should be followed.

For further guidance on the selection of deoiling equipment, the user is referred to the equipment summary
tables presented in section 6.1 and the detailed discussions covering each equipment design presented in
chapters 7 and 8.

6.2.8. Treatment or disposal of secondary streams (8)


All water treatment systems will generate secondary streams such as concentrated hydrocarbon streams,
sludges and stripping streams. Appropriate treatment or disposal methods for secondary streams must be
clearly identified at the design stage with full consideration given to potential Health, Safety and Environmen-
tal implications. This is particularly important for tertiary treatment systems which may involve the removal of
volatile organic compounds such as benzene from water. For more information reference should be made to
section 4.5.

6.2.9. System optimisation and integration (9)


Once the location and basic elements of the water treatment system have been defined, the design should be
reviewed to identify scope for optimising the performance of the system and to ensure the design has been
effectively integrated into the overall production process scheme. System optimisation and integration are
discussed in more detail in section 6.3.

72 Deoiling Manual V 2.0


Restricted

Table 6.2 Selection criteria for water treatment systems

6.3. System optlmlsatlon and integration considerations


6.3.1 Introduction
To optimise the design or operation of water treatment equipment it should be considered as part of an overall
production process, not as isolated, “end of pipe” equipment.

During the design or for the operation of water treatment systems, each of the topics addressed in the fol-
lowing discussions should be considered to ensure the performance of the water treatment system has been
optimised and that the water treatment system is effectively integrated into the overall process scheme.

Deoiling Manual V 2.0 73


Restricted

6.3.2. Maximising droplet size


As the performance of most deoiling equipment is strongly influenced by the size of the dispersed hydrocar-
bon droplet, every effort should be made to maximise the droplet size.
In particular, shear forces which generate small droplets should be minimized or avoided. This is the more
important because the coalescence of small droplets at low concentrations will be minimal, so the droplet
size distribution will not recover in downstream transport.

There are indications that breaking up high pressure drops over multiple valves has a positive effect. Low
shear valves, which are in the final stage of development, have demonstrated significant potential in this
respect [1]. Where possible, control valves should be located downstream of water treating equipment.

Feed pumps should be selected on the basis of having low shear characteristics. Piping routings and fittings
upstream of water treatment systems should be designed to avoid introducing unnecessary shear forces or if
possible located downstream of the water treatment system.

Coalescence of dispersed hydrocarbon droplets should be promoted wherever possible.

6.3.4. Production separators


The main function of most production separators is the separation of gas from liquids and water from hydro-
carbons. While the removal of the water remaining in the hydrocarbon stream must be the primary concern,
consideration should also be given to the minimisation of the hydrocarbons remaining in the water.

Where process conditions are suitable, the use of plate packs or coalescing devices in the production sepa-
rators should be considered to reduce the quantity of hydrocarbons remaining in the water stream.

6.3.5. Stable feed streams


Process equipment operates most efficiently when supplied with a constant, stable feed stream. However
many water treatment systems are located at the end of the process and are subject to wide variations in
process conditions from upstream operations.

Where possible the flow to water treatment equipment should be smoothed by appropriate buffer capacity
on feed streams or through improved control of upstream operations. As well as smoothing flow variations,
the provision of buffer capacity allows for the averaging of peaks in hydrocarbon concentrations and can be
used to promote additional coalescence of dispersed hydrocarbon droplets.

An example is the water stream from continuous or batch dehydration. Water from batch dehydration is often
difficult to treat as the flow rate will vary with the falling head in the dehydration tank, while the hydrocarbon
concentration will increase as the oil interface is approached, on some occasions passing slugs of hydrocar-
bons to the downstream equipment. In contrast, continuous dehydration can result in a water stream which is
stable in both flow rate and concentration, leading to improved deoiling performance.

6.3.6. Recycle streams


The effect of recycle streams on process integration should be carefully considered. Recycle streams may
often be contaminated with solids, chemicals or bacteria and recycling these streams may result in accu-
mulation in the process. In particular, sludges or tight emulsions which accumulate at hydrocarbon/water
interfaces should not be recycled to the process without appropriate treatment to break down the inherent
stability of the sludge or emulsion.

Where possible, reject streams from any equipment should be disposed by injecting into the hydrocarbon
export stream provided sufficient dilution exists to avoid exceeding export quality specifications. This avoids
the use of recycle streams. Alternatively, recycle streams may be processed in dedicated sludge or emulsion
treating facilities.

74 Deoiling Manual V 2.0


Restricted

Treatment requirements for reject streams should be considered as part of the overall water treatment system.
For example, conventional dual media filtration systems require additional equipment for the treatment of dirty
backwash water.

6.3.7. Mixing of water streams


Most hydrocarbon facilities will produce a number of waste water streams from different sources. Mixing of
these streams for treatment in common water treatment facilities may lead to problems such as;
■■ Fluctuating feed rates and contaminant concentrations due to water arriving from several sources. This may
lead to unstable operating conditions in water treatment equipment which will be detrimental to its perform-
ance.
■■ Water from different sources may react to form undesirable compounds or precipitates.
■■ Introducing oxygenated water may result in corrosion problems and the precipitation of oxides from the
water.

To avoid these potential problems, wherever possible consideration should be given to segregation of treat-
ment facilities.

6.3.8. Treatment chemicals


A wide range of treatment chemicals are commonly used to assist the production process such as demulsifiers,
deoilers, antifoams, corrosion inhibitors and hydrate inhibitors.

Whenever a process chemical is used, its potential impact on the entire process system should be consid-
ered. This should include the effect on downstream equipment, interaction with other treatment or residual
chemicals and possible influence on upstream systems through recycle streams.

Examples where the use of treatment chemicals can affect the overall process system include:
■■ Demulsifiersused to improve the dewatering of water-in-oil emulsions can result in the stabilisation of hydro-
carbons droplets in the water phase
■■ Surface active chemicals such as corrosion inhibitors stabilising oil-in-water emulsions.

In addition to potential operating problems, the use of chemicals incurs a range of additional costs such as
the purchase cost of the chemical itself, the cost of transportation, storage and handling of the chemical and
the operating and maintenance cost of chemical injection facilities.

Finally, in many cases, treatment chemicals represent an additional contaminant which may need to be sub-
sequently removed to allow the safe discharge of waste streams into the environment.

Due to these potential drawbacks, it is recommended that wherever possible, process systems should be
designed to operate without the use of corrosion inhibitors and efforts be made to reduce the use of treatment
chemicals wherever possible.

6.4. Examples of water treatment schemes


6.4.1. Draugen (Norway)
Draugen is an oil platform in the North Sea, offshore Norway, processing a light 38° API crude. Fig. 6.2
shows the Draugen oil/water separation process.The produced water treatment plant consists of four par-
allel US Filter hydrocyclone units. The total water handling capacity is 40,000 m3/d. Treated water from
the hydrocyclones flows through pressure reduction valves into a horizontal Produced Water Degassing
Vessel and is then disposed to sea via a caisson. Draugen has also tested Produced Water Reinjection.

The oiw concentration downstream of the hydrocyclones is about 30 ppmw, indicating that they remove all
oil droplets larger than 6 micron. This is in line with the claims of the vendor. The Degassing Vessel plays an

Deoiling Manual V 2.0 75


Restricted

Figure 6.2. Draugen oil water separation process.

important role in the further polishing of the water quality. It was retrofitted with a dedicated sparger and plate
pack to maximize further deoiling. The function of the sparger is to distribute the water and the flashing gas
across the horizontal cross section of the vessel (The branches of the sparger were closed at the ends and
provided with a large number of equally distributed holes of 20 mm diameter at the 12 o’clock position). The
function of the plate pack is to assist with capturing oil and degassing of the froth. Details of the Degassing
Vessel are shown in Fig. 6.3. The overboard oil in water concentration achieved is 15-20 ppmw.

Figure 6.3. Draugen degassing vessel with sparger inlet device.

76 Deoiling Manual V 2.0


Restricted

The recovered oil stream from the hydrocyclone units is routed to the Second Stage Separator. The gas from
the Produced Water Flash Drum is routed to the LP flare system.

6.4.2. Gamba (Gabon)


Fig. 6.4 shows a scheme of the oil/water separation facilities at the Gamba terminal in Gabon. Originally
the bulk oil/water separation occurred in a dehydration tank, followed by desalination in a second, similar
tank as indicated in the scheme. In the current set up the two tanks are both operating as primary separators
in parallell.

Figure 6.4: Gamba terminal oil water processing facilities.

The secondary water treating at Gamba consists of an l’Eau Claire induced static flotation (ISF) unit, see par
7.9. This recovers oil from the produced water from the dehydration tanks (oil content: 150-250 ppm) and
wash water from the Rabi desalination tank (oil content: ~200 ppm average). The recovered oil is pumped
back into the Gamba production header to the dehydration train.

The ISF is an induced gas system in which oil and suspended solids are removed from the oily water by
streams of fine gas bubbles rising up through the vessel and lifting contaminants to the liquid surface. The
ISF is a horizontal cylindrical pressure vessel operating at 0.16 barg. Fuel gas is introduced into the vapour
space as blanketing gas to provide a medium for induced flotation and to protect the vessel against vacuum.
Oily water streams enter the vessel through the inlet chamber and pass successively through each of the four
flotation chambers. A re-circulating flow of blanket gas is induced from the vessel’s vapour space into the oily
water by a recycle stream of clarified water passing at high velocity through eductor nozzles at the bottom
of the vessel.

Eductor supply pumps provide clean water re-circulation. Since the length of each flotation chamber is
approximately equal to the vessel diameter, gas bubbles are equally distributed throughout the oily water.
Floated oil and solids are skimmed from the liquid surface by a simple collection trough. The recovered oil

Deoiling Manual V 2.0 77


Restricted

is collected in a separate chamber of the vessel and is periodically evacuated by skim pumps to one of the
the dehydration tanks. The de-oiled salty water is disposed of to sea. The residual oil in water concentration
is about 30 ppm.

6.4.3. NAM K2/G16 (Netherlands)


Fig. 6.5 shows a typical process scheme of a NAM platform in the North Sea. Key is that the primary con-
densate/water separation is performed in the high pressure part of the system, i.e. without subjecting the
water condensate mixture to the high mixing intensity which accompanies depressurization from the operating
pressure of the production separator to ambient. The water condensate separator is provided with a cross
flow plate pack. The spontaneously occurring dissolved gas flotation in the flash vessel helps to further deoil
the produced water.

Fig. 6.5 Typical North Sea condensate water separation process.

6.4.4. Malampaya (Philippines)


The Malampaya Gas platform is located in shallow water offshore the Philippines). The platform is in produc-
tion since 2001. Methanol is used in the upstream transport as hydrate inhibitor. The water treatment facilities
consist of a methanol feed drum (bulk separator), two parallel coalescers, a methanol recovery tower and
absorption filters. See Figure 6.6 (next page).

From the feed drum the water/methanol mixture is passed through coalescers to remove the last traces of
condensate. From the coalescers the mixture is sent to the Methanol Recovery Tower to recover the methanol.
The bottom stream of the Methanol Recovery Tower is treated in the Produced Water Treatment Package
(PWTP) prior to discharge to the sea. The PWTP has been in operation since July 2005. The PWTP consists
of cartridge filters and two adsorption beds, and is designed to further reduce the oil and phenol concentra-
tion in the produced water.

The system is currently being debottlenecked to achieve design throughput. The adsorber is overloaded due
to incidental breakthrough of condensate through the Methanol Tower, and rapid depletion of the active
components from the adsorber beds, due to insufficient methanol removal. Changes to the operation and
internals of the Tower have been proposed to resolve this. Furthermore the functionality of the somewhat un-
dersized coalescers is to be improved by the retrofit of a platepack.

78 Deoiling Manual V 2.0


Restricted

Fig. 6.6. Malampaya methanol recovery and produced water treating process.

6.5 References
1. Jernsletten, J. ‘Typhoon valve close out’, Report EP 2010-5365.

Deoiling Manual V 2.0 79


Restricted

80 Deoiling Manual V 2.0


Restricted

Chapter 7.

Deoiling equipment –
dispersed hydrocarbons

Deoiling Manual V 2.0 81


Restricted

7. Dispersed hydrocarbons
7.1. Introduction
chapter 7 presents discussions of proven equipment for the removal of dispersed hydrocarbons from water.
Where applicable each discussion includes information on equipment design principles, performance char-
acteristics and sizing guidelines. For the removal of dissolved hydrocarbons from water the user is referred
to Chapter 8.

When considering any process design or operational deoiling problem it is strongly recommended to review
the system design information presented in Chapter 6. Improvements in deoiling equipment design or opera-
tion can often be made by consideration of the deoiling system within the context of the overall process. This
is particularly true with respect to the influence of upstream handling on the droplet size distribution of the
dispersed hydrocarbons entering the deoiling process.

Effective integration of the water treatment system into the overall system design’will result in the most efficient
and cost effective solution.

Following section 7.2 (Definitions), each equipment class is presented in an individual section containing
relevant information for that equipment.

Table 7.1.1 presented on the following pages summarises the general performance characteristics of all the
deoiling equipment covered in Chapter 7.

7.2. Definitions
the performance of deoiling equipment is often expressed as an efficiency. The two most common definitions
are hydrocarbon removal efficiency and process efficiency.

Hydrocarbon removal efficiency is a simple percentage efficiency relating the feed and product oil concen-
tration. It does not consider the quantity of the product and reject streams, only the quality of the product
compared to the feed.
CP
Hydrocarbon removal efficiency (%): E oil = (1 − )x100  
CF
The process efficiency takes the process flow rates into consideration and will thus indicate a poor efficiency
for a process which generates a large reject stream.

Process efficiency (%): €  

Where:
R = reject ratio = QR/QF
CF = oil in water concentration in feed (mgll)
CP = oil in water concentration in product (mg/l)
QF = feed flow rate (m3/h)
QP = product flow rate (m3/h)
QR = reject oil stream flow rate (m3/h)

The major deficiency of these definitions is that they do not consider the size distribution of the dispersed
hydrocarbons in the feed stream. Thus, while a particular class of deoiling equipment may have 100% hy-
drocarbon removal efficiency for 50 µm droplets, it may have a 0% hydrocarbon removal efficiency for 5
µm droplets. Thus these efficiencies are only a measure of deoiling performance at a particular droplet size
distribution.

82 Deoiling Manual V 2.0


Restricted

Aperformancecriterionthatgivesabetterreflectionofremovalefficiencyinrelationtodropletsizedistribu-
tionisthecut-offdropletsize,thisisthesizeofthesmallestoildropletwhichcanberemovedfromthefeed.
Thecut-offdropletsizecanberelatedtotheresidualoilconcentrationwhenthefeeddropletsizedistribution
isknown.Ifthisisnotavailableanestimatecanbemadeonthebasisofanempiricalcorrelation.Figure
7.2.1showsmeasuredoilinwaterconcentrationsattheoutletofvariouswatertreatingequipmentplotted
againstthecorrespondingdropletcutoffsize.Thesedatashowacorrelationoftheform:

concentration(ppmv)≈(dco(micron))1.9forconcentrationsbelow1%v. (7.1)


Fig. 7.2.1. Oil in water concentration at the outlet of deoiling equipment versus droplet cut off size.

DeoilingManualV2.0 83
Restricted

7.3. Coalescers
The performance of most deoiling equipment for the removal of dispersed hydrocarbons is strongly influenced
by the size of the hydrocarbon droplets in the feed stream with larger droplets being easier to separate.
Coalescers are devices designed to promote the coalescence of small hydrocarbon droplets into larger
droplets. Coalescers do not actually separate hydrocarbons from water, however they aim to improve the
separation performance of downstream equipment by increasing the hydrocarbon droplet size in the feed to
these downstream units.

Guidelines for the selection and design of coalescers are dealt with in DEP 31.22.05.12 (liquid/liquid
separators) [1].

84 Deoiling Manual V 2.0


Restricted

7.4. Skimming tanks and vessels


7.4.1. Introduction
Skimming tanks or vessels (collectively described as “skimmers”) are simple gravity based separation equip-
ment, designed for the continuous separation of hydrocarbon droplets from water. Skimmers are often used to
provide surge capacity, ensuring a relatively stable flow to more sensitive downstream separation equipment
and allowing provision for the removal of large slugs of oil which may ovewhelm downstream equipment.

Skimmers can be either vertical or horizontal designs. In vertical skimmers the dispersed hydrocarbon droplets
must rise upwards against the downward water flow while in horizontal skimmers the dispersed hydrocarbon
droplets rise perpendicular to the horizontal water flow.

Figure 7.4.1 illustrates a simple vertical skimming tank. The water stream leaves the skimmer via the dis-
charge ring main. Hydrocarbon droplets above a critical diameter will rise against the downward water flow
and can be skimmed from the water surface. Due to the large cross sectional area the skimmer can be readily
designed to accommodate degassing and solids removal.

In horizontal skimmers the vertical rising of the hydrocarbon droplets is superimposed over the horizontal flow
of the bulk water flow. The length of the horizontal flow path must be sufficient to allow the hydrocarbon
droplets to reach the hydrocarbon layer on the surface of the water.

Figure 7.4.1: Schematic diagram of simple skim tank.

7.4.2. Performance variables


The performance of a skimmer is generally governed by the variables composing Stokes law. Stokes law is
discussed in detail in Appendix A2.1.

As with all separation equipment governed by Stokes law, the hydrocarbon droplet size is the most important
variable affecting the separation performance. Droplet sizes should be maximised by avoiding shearing and
promoting coalescence wherever possible.

Skimmers are usually sized for separation of hydrocarbon droplets of size 150 µm or larger.

7.4.3. Configuration
The configuration of the internals in the skim tank/vessel will also have a significant effect on the separation
performance. In a recent optimization study of a skim tank for PDO the separation efficiency of a number of
different configurations was compared by means of CFD particle tracking calculations.[1] It could be estab-
lished that configurations with tangential inlets were more efficient than the ones with radial vane type inlets.
See Figures 7.4.2 and 7.4.3.

Deoiling Manual V 2.0 85


Restricted

Fig. 7.4.2. Alternative inlet and outlet configurations ND 1-3. The base case (original configuration) was similar to ND1, with the inlets separated 90°
instead of 180° .
Efficiency

Fig. 7.4.3. Separation efficiencies for the different configurations of Fig. 7.4.2.

86 DeoilingManualV2.0
Restricted

7.5 Discharge caissons


7.5.1. Introduction
Most offshore platforms discharge waste water through a caisson suspended from the platform and extend-
ing below the water surface. Although primarily a conduit for water disposal, the caisson can also serve the
following functions.
The caisson can be sized to allow the separation of large hydrocarbon droplets from water. Although the
water zone reporting to the caisson should have been treated to the required environmental discharge quality
requirements, the caisson provides a location to capture any accidental zone hydrocarbon discharges before
they enter the environment.
The caisson can provide an alarm or shut-down point in the event of accidental hydrocarbon discharge to
the caisson. The disposal caisson must not be considered as an integral part of the water treatment system.
Its function is to act as a back-up device which will help minimise the potential for any large accidental hy-
drocarbon discharge in the event of failure of upstream equipment.

At the time of writing there is no detailed information available on the process design or operation of cais-
sons.

7.5.2. Design guidelines


Caissons typically have an inside diameter in the order of 600
to 1200 mm. If possible the diameter should be set on the basis
of Stokes law to allow a hydrocarbon droplet of 150 mm to rise
against the downward flowing water.

This may not be practical for caissons with high discharge flow
rates such as when disposing of large quantities of produced
Skim pile water and in many cases the diameter of the caisson
may be set by structural considerations.

7.5.3. Caisson internals


lnternals can be fitted to improve the hydrocarbon separation
performance of a disposal caisson. However, with the increased
complexity of upstream deoiling equipment (e.g. hydrocyclones
followed by degassing tanks capable of capturing hydrocarbon
slugs), the use of caissons incorporating internals and oil recov-
ery systems has declined.

Skim piles
The skim pile design is illustrated in Figure 7.5.1. The design
provides a series of baffles which assist the separation of hydro-
carbon by:
■■ Providing quiescent areas under the baffles with minimum net
downward water velocity. This allows small droplets to be col-
lected.
■■ The plates minimise the actual distance that a droplet has to
rise before it will intercept a surface and be collected.

SP Piles
MPE market a caisson design fitted with a number of SP Pack
coalescing packs to improve the separation performance. Figure 7.5.1: Skim pile.

Deoiling Manual V 2.0 87


Restricted

Closed caissons
Caissons can be designed with a closed bottom with the level inside the caisson controlled by a siphon leg.
A closed caisson reduces the effects of wave motion and maximises the water residence time in the caisson,
allowing for better removal of accidentally discharged hydrocarbons. Provision would have to be made for
the removal of accumulated solids from the bottom of the caisson.

7.5.4. Operational considerations


A caisson will only operate satisfactorily during periods of relatively low wave action. As wave height grows,
the caisson will be subject to a piston pumping action and will be unlikely to achieve any significant separa-
tion.

88 Deoiling Manual V 2.0


Restricted

7.6. API separator

API separators are gravity type separators for onshore installations, designed according to principles estab-
lished by the American Petroleum Institute (API). Although the design has been largely superseded by alter-
native designs (particularly the Corrugated Plate Interceptor (CPI)), there are a large number of existing API
interceptors remaining in use within the petroleum industry.

Therefore the API separator is briefly discussed in Appendix A7.1.

Deoiling Manual V 2.0 89


Restricted

7.7. Plate interceptors


7.7.1. Introduction
A plate interceptor consists of a pack of inclined parallel plates with the oily water flowing through the chan-
nels created between the plates. The close proximity of the plates ensures each flow channel has a relatively
small effective hydraulic radius. This maintains stable laminar flow conditions through the plate pack at a
higher flow rate than for an empty vessel and allows small hydrocarbon droplets to rise relatively undisturbed
by turbulence.

Within a plate pack, a hydrocarbon droplet will have a forward velocity corresponding to the net water
flow and a rising velocity as defined by Stokes law. The proximity of the plates ensures that the droplet only
has a small distance to rise before it is intercepted by the underside of the upper plate. The collection of
hydrocarbon droplets on the underside of the plate promotes coalescence and the large coalesced droplets
can rise along the plate for collection.

Common alternative names for various designs of plate interceptors are Parallel Plate Interceptor (PPI), Tilted
Plate Interceptor (TPI) and Corrugated Plate Interceptor (CPI). A Shell standard CPI design is available and is
commonly used for onshore installations.

Figures 7.7.4 to 7.7.6 on the next pages illustrate a range of possible configurations.

7.7.2. Configurations
There are three basic plate pack flow configurations, cross flow, downflow and upflow. Cross flow and
downflow are the most common. The Shell design practice is based on cross flow.

Cross flow
A cross flow design is illustrated in Figure 7.7.1. The
water flows horizontally across the plates and does
not preferentially favour the separation of hydrocar-
bons or solids.
The cross flow design has the advantage in that it can
be efficiently packaged into different configurations.
As it does not need excess space above or below the
plates for the entry and exit of the water flow it can
often utilise more of the available vessel volume than
alternative designs. This is a particular advantage in
pressure vessels, where a cross flow design can fill
more of the available cross-sectional area with plates Figure 7.7.1
Cross flow plate geometry.
and thus minimise the overall size of the vessel.

Down flow
A down flow design is illustrated in Figure 7.7.2. The advantage
of the down flow design is that the downwards flow through the
pack will help sweep solids from the plate pack, minimising the
potential for blockage.

The disadvantage of a down flow design is that the downwards


flow through the pack may disturb the upwards flowing hydrocar-
bon film on the underside of the plates. Hydrocarbons which are
disturbed in this manner will remain in the clean water stream. Figure 7.7.2
Down flow plate geometry.

■■ In the absence of other factors, the cross flow design is a good compromise offering efficient packaging,
a horizontal flow path and neutral behaviour towards both solids and hydrocarbon droplets.

90 Deoiling Manual V 2.0


Restricted

■■ The horizontal flow path of the cross flow design often makes it the simplest design to retrofit to horizontal
separation equipment.
■■ When a high solids load is expected the down flow configuration is often the best choice as the down-
wards flow keeps the solids moving and prevents blockage of the plates.

Corrugated plates
In a plate pack consisting of flat plates, the hydrocarbon droplets accumulating on the bottom surface of the
plates will be relatively evenly distributed across the surface of the plates. This even distribution will not assist
coalescence of the hydrocarbon droplets.

Figure 7.7.3 illustrates the behaviour of hydrocarbons and sol-


ids in a down flow corrugated plate pack with the flow along
the corrugations. A cross flow configuration would typically
have the flow entering from the side, flowing across, not along,
the corrugations.

The rising hydrocarbon droplets will be concentrated in the con-


cave surfaces under the corrugated plates. This accumulation in
the corrugation will assist in the coalescence of the droples and
improves the separation performance of the plate pack. Solids Figure 7.7.3
will tend to concentrate in the troughs on the upper surfaces of Corrugated plates.

the corrugated plates.

Figure 7.7.4
Multiphase separator incorporating plate pack (Skimovex).

Deoiling Manual V 2.0 91


Restricted

Figure 7.7.5
Down flow plate pack – possible pressure vessel configuration (Skimovex).

Figure 7.7.6
Cross flow plate pack – pressure vessel configuration (Rossmark-Pielkenrood).

92 Deoiling Manual V 2.0


Restricted

Unfortunately, no information has been found which can actually demonstrate whether corrugated plates have
better coalescing properties than flat plates. On this basis both flat plates and corrugated plates must be
considered equally acceptable. The only constraints on the use of corrugated plates are;

If solids are present the corrugations must be aligned to allow solids to move out of the plate pack without
being trapped by the corrugations.

7.7.3. Performance and sizing of plate packs


The cut-off droplet size achievable with a plate pack separator is determined by the flow velocity of the con-
tinuous (water) phase, the plate spacing and the plate length, and can be estimated on the basis of Stokes
law. The size of the smallest droplet that can be separated will be of the order of 50-60 microns.
Ideally the channel flow between the plates should be laminar. Therefore the Reynolds number should be
limited to 850 for flat plates, and 450 for corrugated plates. In many cases, e.g. in retrofits, this will not be
achievable but there is evidence that a plate pack will improve the separation, even at Reynolds numbers as
high as 5000.
The design of separators with plate packs is dealt with in more detail in DEP 31.22.05.12.

7.7.4. Installation
Plate interceptors can be built into pressure vessels and thus can be installed at any point in the water treat-
ment process. They are often installed as a first stage in a water treatment system, designed for removal of
the bulk of the dispersed hydrocarbon before a second treatment stage. However where the process allows
the plate pack should be used as far upstream as possible, upstream of level control valves and if conditions
permit, even in primary separators.
Plate packs are often successfully used for single stage treatment of streams with large oil droplets such as
deck and drains water and ballast water treatment.

7.7.5. Operational considerations


Solids handling
If solids are expected in the feed stream the design should allow for their removal. Rapidly settling solids can
be allowed to settle in a stilling zone before entering the plate pack. Finer solids and sludge will settle at
the bottom of the plated section. Appropriate nozzles and jetting systems should be provided for removal of
accumulated solids.

Plate pack removal


In many applications the plate pack may eventually be blocked due to fouling or accumulation of solids.
Provision should be made for suitable access and lifting equipment to allow the removal of the plate pack for
cleaning or replacement as required.

7.7.6. Concrete based standard CPI design


Shell has a standard design for an open, below ground, concrete based, down flow corrugated plate
interceptor as illustrated in Figure 7.7.7. This design is commonly used for land based installations. The
design of such a CPI is covered in detail in the following documents.
■■ “Drainage and primary treatment systems”, DEP 34.14.20.31-Gen, 2010, and accompanying Standard
Drawings
■■ “Deoiling industrial waste water. Design and operation of the “CPI” (Corrugated plate interceptor)”, Report
MF 81-8700, May 1981.
These documents include example calculations, standard dimensions, and cover all civil and construction
requirements for the CPI design. It should be noted that the normal capacity of the Shell standard design is
based on a Reynolds number of 400.

Deoiling Manual V 2.0 93


Restricted

Figure 7.7.7
Illustrative schematic of standard Shell CPI design.

94 Deoiling Manual V 2.0


Restricted

7.8. Hydrocyclone
7.8.1. Introduction
The hydrocyclone has rapidly been accepted as a compact and efficient means of removing dispersed hy-
drocarbons from water. The basic design is illustrated schematically in Figure 7.8.1.

Figure 7.8.1
Schematic diagram of a hydrocyclone.

The water containing the dispersed hydrocarbons enters the hydrocyclone through a tangential inlet at the top
of the swirl chamber. As the liquids swirl along the hydrocyclone, the centrifugal forces generated promote
the separation of the hydrocarbon and water phases, with the hydrocarbon phase forming a thin core at the
centre of the hydrocyclone.

By maintaining a suitable pressure ratio between the clean water outlet stream and the reject oil outlet stream,
the geometry of the hydrocyclone will result in the thin hydrocarbon core flowing in a reverse direction, exiting
from the top of the swirl chamber. The clean water exits from the tail section of the hydrocyclone.

7.8.2. Definitions
The following definitions are commonly used:
Feed: The oily water feed stream entering the hydrocyclone
Underflow: The clean water stream exiting from the tail of the hydrocyclone
Reject stream or Overflow: The concentrated hydrocarbon stream exiting from the head of the
hydrocyclone though the reject port.
Reject ratio (R): The ratio of the reject and feed stream volumetric flow rates: R=Qreject/Qfeed

7.8.3. Performance variables


The flow patterns and forces acting within a hydrocyclone are very complex and unlike traditional gravity
based separation equipment cannot be simply described in terms of settling velocity.
However, in general terms, the variables present in Stokes law will have a strong influence on the separation
performance of the hydrocyclone and these variables are examined in detail in the following discussions.

Deoiling Manual V 2.0 95


Restricted

Hydrocarbon droplet size


Stokes law (Appendix A2.1) illustrates that diameter is the dominant variable affecting the buoyancy of a
hydrocarbon droplet. Although Stokes law cannot be directly applied to a hydrocyclone, the droplet size will
have a similar effect on the buoyancy forces which cause the hydrocarbon droplet to migrate to the centre of
the hydrocyclone. Thus, as the size of a hydrocarbon droplet increases, the rising velocity of the droplet in-
creases and the droplet has a higher probability of migrating to the reject oil stream. Similarly, as the droplet
size decreases, there will be a reduced probability that the droplet will migrate to the reject oil stream and
an increased probability of the hydrocarbon droplet exiting with the water stream.

As a result, it is important to maximise the droplet size reaching the hydrocyclone by avoiding shearing and
promoting coalescence wherever possible. Hydrocyclones should be located close to the source of the oily
water, preferably upstream of control valves. For pumped hydrocyclone installations, the feed pumps should
be a low shear design.

Separation of droplet sizes in the order of 6 to 7 µm can be achieved using state of the art hydrocyclone
designs (e.g. Vortoil K liner, US Filter) if conditions are favourable (e.g. high density difference between the
phases, low water viscosity). This corresponds to a residual oil in water concentration of about 30 ppm, cf.
Figure 7.2.1.

Flow rate
The physical geometry of the hydrocyclone is designed to generate the optimum swirl characteristics at a par-
ticular design flow rate, maximising the centrifugal acceleration while minimising shearing of the dispersed
phase into smaller droplets.

At low flow rates, separation is reduced due to lower centrifugal forces. In general, to achieve satisfactory
separation it is recommended that hydrocyclones are not operated below 40% of their design capacity.
At some point above the design flow capacity the performance of the hydrocyclone will deteriorate. This
deterioration can be due to a number of factors. At higher flow rates re-entrainment may be experienced
in the hydrocyclone, reducing separation performance. As flow rates increase, the pressure drop across
the hydrocyclone will increase until eventually the available feed pressure is not sufficient to drive the reject
stream out of the hydrocyclone. Even before this point is reached, the increased pressure drop may lead to
gas breakout in the hydrocarbon core, upsetting the performance of the hydrocyclone.

Reject stream ratio


The discharge rate of the reject stream must be adjusted to ensure that the central hydrocarbon core is re-
moved from the hydrocyclone. Below a minimum reject ratio, the efficiency of the hydrocyclone will be poor
as the reject stream flow rate will be too low to purge the hydrocarbons from the hydrocyclone and some
hydrocarbons will be discharged with the water stream.

Figure 7.8.2 illustrates the typical relationship between oil removal efficiency and reject ratio. As the reject
ratio is increased, the efficiency of the hydrocyclone improves. Higher reject ratios will achieve marginal
improvements in hydrocarbon removal efficiency, however this is at the expense of a larger volumetric flow
rate of reject stream, hence a lower process efficiency.
The optimum reject ratio will be dependent on the particular operating conditions of the hydrocyclone instal-
lation. Satisfactory efficiencies have been obtained with reject ratios as low as 0.5%, especially with the
assistance of mitigating factors such as large density differences between the hydrocarbon and water phases
or elevated temperatures to reduce the water viscosity.

Although operation at the minimum reject ratio has the advantage of minimising the flow rate of the reject
stream, in practice an operating margin should be provided above the minimum reject ratio. This margin as-
sists the hydrocyclone to maintain satisfactory performance under upset conditions. A reject ratio in the order
of 1.5 -2.0% is typical.

96 Deoiling Manual V 2.0


Restricted

Figure 7.8.2
Variation in hydrocyclone hydrocarbon removal efficiency with reject ratio.

During severe upset conditions, with the presence of hydrocarbon inlet concentrations greater than 1% (wt)
(10,000 mg/l), the reject ratio may be increased to assist in the removal of the excess hydrocarbons. The
required reject ratio for operating under upset conditions should be determined from operating experience.
As a starting point, 2% per % hydrocarbon content in the feed may be used.

Density difference
The performance of a hydrocyclone will improve with increasing density difference. Thus in principle a light
condensate should be easier to separate than a heavy oil.

It should be remembered that characteristics of the hydrocarbons other than density difference will also
have an impact on the separation performance. Thus a condensate with a high density difference may be
considered easy to separate, however other factors such as the ability of the condensate to coalesce will also
have a significant effect on the overall separation performance.

Temperature/Viscosity
Variations in temperature will influence both the density difference between the phases and the viscosity of
the continuous water phase. In general, higher temperatures assist separation, particularly by reducing the
viscosity of the water phase. However, in some cases, adverse density changes with temperature can hinder
the phase separation.

The influence of the water viscosity is illustrated in Figure 7.8.3. The effect is presented in terms of d75,
where d75 is a droplet which has a 75% chance of being removed. d75 is closely related to the Stokes cut
off droplet size used elsewhere. It can be seen that d75 rises with increasing viscosity, indicating reduced
performance. Thus placing the hydrocyclones in a location where the water temperature is maximised will
assist in maintaining a high separation efficiency.

Temperature may have additional effects on the hydrocarbon/water dispersion, such as altering surface ten-
sion or altering dispersion stability. These effects may influence the overall hydrocyclone performance.

Deoiling Manual V 2.0 97


Restricted

Figure 7.8.3
Variation in hydrocyclone performance with continuous phase viscosity.

Pressure drop - Definitions


Under normal operation there are two distinct pressure drops across a hydrocyclone.

DP1 = Pin – Pout

DP2 = Pin - Prej

Where: Pin = Feed stream pressure


Pout = Water outlet stream pressure
Prej = Hydrocarbon reject stream outlet

The pressure drop to the reject stream (DP2) is the most significant as it is the greater of the two and will ulti-
mately determine the capacity of the hydrocyclone.

C, the ratio between DP2 and DP1, has a linear relationship with the reject ratio, as illustrated in Equation
7.8.1.

  Eq. 7.8.1

where C = constant 

This relationship is schematically illustrated in Figure 7.8.4. The slope of the curves in Figure 7.8.4 are char-
acteristic for each hydrocyclone installation, but are mainly a function of the geometry of the hydrocyclone.
As the pressures in and out of the hydrocyclone are easily measured, this relationship may be used as the
basis for measurement and control of the reject ratio of the hydrocyclone.

Minimum pressure drop


The minimum allowable pressure drop (inlet to reject stream) across a standard hydrocyclone is typically of
the order of 4 bar. Dedicated hydrocyclones for low pressure drop applications have minimum pressure drops
as low as 2 bar.

98 Deoiling Manual V 2.0


Restricted

Figure 7.8.4
Relationship between reject ratio and C.

Both the capacity and the turn-down characteristics of a hydrocyclone improve with increasing pressure drop.
Use of a feed pump to boost the hydrocyclone feed pressure and hydrocyclone capacity may be a more
cost effective solution than a greater number of lower capacity hydrocyclones operating from a low process
pressure. However the feed pump must be a low shear design to avoid additional shearing of the dispersed
droplets.

Maximum allowable pressure drop


The maximum allowable pressure drop across a hydrocyclone may be limited by a number of factors.
■■ Erosion considerations, especially if solid particles are present in the water stream.
■■ Excessive shear energy input, resulting in reduced separation efficiencies.
■■ Gas breakout in the hydrocarbon core due to the reduced pressure, interfering with the separation process.

The hydrocyclone vendor should advise the maximum recommended pressure drop for a given hydrocyclone
and process configuration. As a first estimate, pressure drop should be limited to a maximum of 30 bar.

Gas breakout
The effect of gas breakout on hydrocyclone performance should be considered. High pressure drops across
hydrocyclone systems can result in the formation of gas bubbles. Gas bubbles may impose additional shear
forces on dispersed hydrocarbon droplets and may upset the stability of the hydrocarbon core at the centre
of the cyclone.

The performance of hydrocyclones will deteriorate when more than a limited fraction of free gas is present
in the hydrocyclone.

7.8.4. Installation/Configuration
Process location
For systems to be operated from process pressure, the hydrocyclone should be preferably be located as close
as possible to the source of the waste water, preferably upstream of level control valves to minimise shearing
of hydrocarbon droplets. Hydrocyclone installations can be readily designed for high pressures (120 bar

Deoiling Manual V 2.0 99


Restricted

designs have been proposed for some gas production installations) and can thus can be installed immediately
downstream of high pressure separators.

With multiple sources of water (e.g. independent separator trains), dedicated hydrocyclone systems for each
source should be given preference to a single hydrocyclone package with individual level control valves
located upstream of the hydrocyclones.

For pumped hydrocyclone systems low shear feed pumps must be specified. These are commonly progressive
cavity pumps or specially designed low shear centrifugal pumps.

Packaging
Depending on the available pressure drop the capacity of a single hydrocyclone is of the order of 4 to
20 m3/h. To achieve the capacity required to process larger flow rates, individual cyclones are assembled
into packages with multiple hydrocyclones in parallel. The most common packaging bundles a number of
hydrocyclones into a pressure vessel with common header space for inlet and outlet streams.

This modular packaging of most hydrocyclone installations has a number of distinct advantages.
■■ The capacity of the water treatment system can be upgraded by adding additional hydrocyclone pack-
ages. Thus future water treatment capacity can be phased in as required.
■■ Turn-down of the system can be accomplished by adjusting the number of hydrocyclone modules on-line.
This ensures on-line hydrocyclones are operating at optimum efficiencies.
■■ It may be possible to upgrade or replace the hydrocyclone liners whilst retaining the external pressure ves-
sel housing.
■■ The possibility of water escaping from the system is reduced as the number of flanges and connections is
minimised and as the hydrocyclone liner is enclosed by a pressure vessel, erosion of the liner will not result
in external leakage.
■■ Modular hydrocyclone packages may be dedicated to individual water sources, thus allowing level control
valves to be located downstream of the hydrocyclones.

Size, Weight and Capacity


The capacity of an individual hydrocyclone is governed by the available pressure drop. Thus the total capac-
ity of a hydrocyclone package is a function of the available pressure drop and the number of hydrocyclones
in the package.

To process flows larger than that handled by a single package, a number of individual packages are com-
bined to give the required capacity. Even when a single package is sufficient to process the total waste
water flow, consideration may be given to a system comprising a number of smaller packages to allow more
efficient operation during turn-down, or to be able to isolate individual hydrocyclones within the package.
Similarly, individual packages for each water source allow the location of level control valves downstream of
the hydrocyclone, minimising droplet shear.

A hydrcocyclone package for 500 m3/h (2 x 50%) has a weight of about 10 tons and a footprint of 6-8 m2.

7.8.5. Operational considerations


Blockage of the reject stream exit port
The size of the hydrocarbon reject stream exit port is generally quite small, in the order of a few millimetres.
Any foreign material which enters the hydrocyclone with a density lower than that of water will migrate to the
centre of the hydrocyclone and may block the reject stream exit port. The effects of such blockage can be
minimised through the use of strainers and/or a backwash system.

Strainers can be specified upstream of the hydrocyclone package to remove foreign materials before they
enter the hydrocyclone. Such strainers are particularly useful during start-up and commissioning. However,

100 Deoiling Manual V 2.0


Restricted

there is a possibility that strainers may induce additional shearing of the hydrocarbon droplets. As such strain-
ers should be selected on the basis of having a relatively coarse aperture size (in the order of 300 µm) and
minimum pressure drop.

A backwashing system should be specified on the reject oil stream. The backwashing system uses a stream
of water or gas to reverse the flow in the reject stream and remove any material blocking the exit port. The
underflow from another hydrocyclone may be used for backwashing.

The reject stream should be blocked in downstream of the backflushing point to ensure the full flow of the
backwashing is directed into the reject port of the hydrocyclone and that any foreign material is carried out
though the underflow of the hydrocyclone. Due to the short residence time in the hydrocyclone backflushing
is only required for a short period e.g. 15 seconds.

Effect of solids
As sand has a density higher than water it will migrate to the walls of the hydrocyclone and exit with the water
stream. Small quantities of sand should not have a direct impact on the performance of the hydrocyclone,
however the sand may have a number of indirect effects such as:
■■ Erosion of the liners.
■■ Potential accumulation of solids in the outer vessel shell of a hydrocyclone packaged module. In some
cases, solids may also potentially block the relatively small hydrocarbon reject port.
■■ Fine solids may stabilise a smaller hydrocarbon droplet, reducing the performance of the hydrocyclone.
■■ Hydrocarbons adhering to the solids will contribute to the hydrocarbons exiting with the water stream.

The allowable sand content in the feed to a hydrocyclone will be dependent on a number of variables that
have a bearing on erosion such as the fluid velocity in the cyclone (pressure drop), the materials of construc-
tion, the hardness of the sand etc. As such it is not possible to simply define a limit on sand content.

Some erosion of liners has been experienced for the Shell Expro North Cormorant hydrocyclones. The Shell
Expro Eider installation uses upstream sand cyclones to reduce the solids loading on the deoiling hydrocy-
clones. An alternative is the use of erosion resistant materials for the hydrocyclone liners such as the use of
boron diffused inconel or Stellite inlet sections combined with stainless steel reducing sections.

When hydrocyclones are bundled together in pressure vessels with a common head collecting the water
discharge, provision should be made for removal of any sand that may collect in the head.

In severe cases, sand will need to be removed upstream of the hydrocyclone system to avoid problems with
blockages or erosion. In this situation consideration should be given to a sand removal system that can also
assist in the coalescence and removal of hydrocarbon droplets.

Operating experience
Hydrocyclones are considered a well established technology with numerous installations throughout the
world.

Deoiling Manual V 2.0 101


Restricted

7.9. Induced gas flotation


7.9 .1. Introduction
Gas flotation is a process by which dispersed hydrocarbon droplets are removed from water by attachment
to rising gas bubbles. The oily froth which forms on the surface of the water is removed by skimming or
overflowing to a collection trough. Gas flotation units are generally installed as a secondary stage in a water
treatment system.

Inert gases, air or hydrocarbon gas are normally used for induced gas flotation. In many cases the use of air
is not practical due to the need to exclude oxygen from the process (the presence of oxygen leads to con-
cerns with corrosion, biological growth and explosion hazard aspects). Chemical additives are often used to
improve the performance of the flotation process.

Gas flotation units are categorised by the method used to generate the bubbles, either Induced Gas Flotation
(IGF) (mechanical and hydraulic) or Dissolved Gas Flotation (DGF). Induced gas flotation is the more com-
mon in E&P operations and is discussed in this section. Dissolved gas flotation is discussed in section 7.10.

Mechanically induced gas flotation


Figure 7.9.1 illustrates the design of a single mechanically induced gas flotation cell. Rotation of the impeller
draws gas down the stand-pipe where it comes into contact with the oily water. The action of the impeller
forces the water and gas through the disperser which results in the formation of small gas bubbles.

The bubbles rise to the surface of the water, collecting hydrocarbon droplets and forming an oily froth on the
surface of the water. The froth flows over a weir into an collecting launder, often assisted by mechanically
driven paddles. The clean water underflows though a port or under a baffle near the bottom of the cell.

Figure 7.9.1: Schematic diagram of a Wemco mechanically induced gas flotation cell. .

102 Deoiling Manual V 2.0


Restricted

Hydraulically induced gas flotation


The operation of a hydraulically induced gas flotation cell is very similar to the mechanically induced gas
flotation cell. However, instead of using an mechanically driven impeller to generate bubbles, a recirculated
stream of clean water is injected into the cell through a nozzle designed to create a venturi effect. Air or
gas is drawn into the cell by the venturi and dispersed as bubbles. The injected water flow also agitates the
fluid in the cell. Hydraulically induced gas flotation is sometimes referred to as Induced Static Flotation (ISF).

Figure 7.9.2 illustrates the design of a hydraulically induced flotation cell which uses top mounted gas educ-
tors. A typical four cell package is shown. Elbow channels are used between the cells to direct the water flow
to the top of each cell and minimise short circuiting across the bottom of the cell.

Figure 7.9.2: Schematic diagram of a four cell Monosep hydraulically induced gas flotation package. .

The advantage of the hydraulically induced design is that it replaces the mechanical driver required for each
cell with a single recirculation pump that may be common for a number of cells. This can also improve reli-
ability as a back-up recirculation pump can be provided. To capitalise on this mechanical simplicity, most
hydraulically induced gas flotation designs also eliminate the mechanically driven skimming paddles by using
alternative weir or level control arrangements.

Hydraulically inducing the bubbles is believed to result in lower shear forces in the flotation cell and a more
complete distribution of bubbles than an equivalent mechanically induced system. In theory, this should make
the hydraulically induced system more effective at capturing dispersed hydrocarbon droplets and more suit-
able for flotation of flocculated hydrocarbon droplets.

Figure 7.9.3 illustrates a cross section of an alternative design for a hydraulically induced gas flotation cell
which operates in a pressure vessel. To maximise the working volume of the pressure vessel the cells use a
central, top mounted froth collection trough and bottom mounted eductors.

A disadvantage of hydraulically induced flotation cells is that the recycled water stream reduces the capacity
of a given cell size.

Deoiling Manual V 2.0 103


Restricted

Figure 7.9.3: Schematic diagram of a L’eau Claire hydraulically induced gas flotation cell. .

7.9.2. Performance variables


Hydrocarbon removal efficiency
The hydrocarbon removal efficiency of a single flotation cell is typically of the order of 60 to 70%. Operating
experience has indicated that the packaging of four cells in series results in an overall hydrocarbon removal
efficiency in the range of 70-99%.

Bench scale flotation equipment can be used on site to investigate the use of chemicals to improve the flota-
tion performance.

Hydrocarbon droplet size


Test work has indicated that the efficiency of flotation is not significantly affected by droplet size, maintaining
relatively high separation efficiencies with droplet sizes in the order of 10 microns. Experience suggests that
the flotation process can remove droplets down to the order of 5 mm.

However, as with most deoiling equipment, performance will improve with larger droplet sizes. Thus it is
important to maximise the droplet size reaching the flotation cells by avoiding shearing and promoting coa-
lescence wherever possible.

Gas input/Bubble size


The quantity of gas entrained in the water for a mechanically induced gas flotation system is typically in the
order of 8.5 m3 gas/m3 water, however most of this gas volume is simply recirculated. The quantity of exter-
nal make up gas required is typically in the order of 0.05 to 0.10 m3 gas/m3 water, sufficient to maintain
a gas blanket above the liquid and compensate for losses. Some operators have experienced higher gas

104 Deoiling Manual V 2.0


Restricted

consumption, in the order of 0.2 to 0.4 m3/m3. The actual mass flow rate of gas consumed must take into
account the operating pressure of the flotation cell. Units operating at higher pressures will require higher gas
mass flow rates to achieve a satisfactory bubble density in the flotation cell.

When using hydrocarbon gas, the gas is often taken from the fuel gas system, through a suitable let down
and overpressure protection system.

Theoretical analysis of the flotation process indicates that the hydrocarbon removal performance of the
flotation cell should be improved by increasing the gas bubble concentration and reducing the gas bubble
diameter. The bubble size for a IGF unit is typically in the order of 100 to 400 mm. Smaller bubble sizes in
the order of 40 to 70 microns are achievable in DGF systems.

The gas bubble diameter is not easily controlled in the field. Higher water salinity and some chemical flotation
aids will help support a smaller bubble size.

Gas bubble concentration is also difficult to control in practice. Up to a point, increasing gas bubble concen-
tration will improve removal efficiency due to an increased probability of bubble/droplet contact. However
further increases in gas concentration will eventually become counter productive by reducing liquid residence
time in the cell and through coalescence of small efficient gas bubbles into larger inefficient gas bubbles.
Large gas bubbles may lead to slugging of the fluids in the cell.

Inlet hydrocarbon concentration


Trial work has indicated that the separation efficiency of a flotation unit is relatively constant over a range of
150 to 2300 mg/l of hydrocarbon in the feed stream.

Very high concentrations of hydrocarbon or slugs of pure hydrocarbon are known to upset the performance
of flotation units, often hindering the establishment of a froth layer. The flotation unit should be protected from
such surges in hydrocarbon concentration by an upstream primary separation stage. Designing the control
system of the flotation unit to maintain a overflow from the surface of the cells under all conditions will assist
in the removal of any accumulated hydrocarbon layer from the water surface.

Flow rate
Figure 7.9.4 illustrates the variation in efficiency of a Wemco four cell gas flotation unit with flow rate. The
fall in efficiency is related to the lower residence time in each cell with increasing flow rate.

The residence time in an induced gas flotation (IGF) system is in the order of four minutes, on a basis of ap-
proximately one minute per cell.

Reject stream
The froth that is skimmed from the surface of the water contains a significant quantity of water. In general, the
reject stream is in the order of 1 to 3% of the total incoming water stream.

Some operators have found it advantageous to operate the flotation unit with a reject stream flow rate in the
order of 5-10% of the incoming water stream to maintain a higher skimming rate and allow the unit to quickly
recover from high hydrocarbon concentrations in the inlet stream.

Chemical addition
In many cases, satisfactory hydrocarbon removal performance can be obtained from gas flotation without the
use of any treatment chemicals. As with all waste water treatment systems it is preferable not to use chemicals
if possible as these will often be discharged with the treated effluent into the environment. However in some
cases chemicals may be required to ensure adequate deoiling performance.

Deoiling Manual V 2.0 105


Restricted

Figure 7.9.4: Realtionship between hydrocarbon removal efficiency and flow rate for induced gas flotation. .

Trial work has indicated that in some cases the performance of gas flotation may be considerably enhanced
by using a suitable flotation aid.

Flotation aids may act through several mechanisms such as promoting the coalescence of the dispersed hy-
drocarbon droplets, enhancing the attraction between the hydrocarbon droplets and the rising gas bubbles
or improving bubble and froth stability. This should be injected well upstream of the flotation cells to ensure
effective mixing and to provide time for the chemical to act. General experience with flotation aids indicate
that there is an optimum dosage rate. At low dosage rates performance is poor, but improves as the injection
rate increases. At excessively high dosage rates the performance falls off again and the chemical can react
with the froth to form a “jelly like” substance that can accumulate on the froth discharge equipment such as
weirs, paddles and launders.
The correct flotation aid and dosage rate can only be selected on the basis of field trials. Typical dosage
rates are in the order of 5 to 50 mg/l.

Suspended solids
The effect of suspended solids on the flotation performance should be carefully considered. In some cases
solids can be effectively floated out of the cells with the froth, allowing the flotation cell to remove both solids
and hydrocarbons. Trials in the USA have claimed 70-99% removal of solids.

However in other cases poor performance of flotation systems has been experienced with water containing
high levels of suspended solids. In these cases the agitation in the flotation cell will carry some solids into the
froth, but will also prevent other solids settling and these will then be carried out with the water stream. Solids
may carry hydrocarbon droplets or hydrocarbon films, contaminating the effluent water.

Hydraulically induced flotation systems use a recirculated water flow and venturi type induction systems to
generate gas bubbles. These components may be subject to erosion if solids are present in the recirculated
water stream.

106 Deoiling Manual V 2.0


Restricted

The flotation of some solids may be improved by the selection of suitable flotation aids. The minerals industry
has a long history of separating solids using appropriate flotation aids to selectively alter the flotation char-
acteristics of solids.

Water salinity
Some test work has indicated that the hydrocarbon removal efficiency improves as salinity increases up to
approximately 3 to 4%. After this point the efficiency remains constant. The mechanism of this improvement
is probably related to a number of factors such as the presence of the electrolyte ions altering the surface
characteristics of the bubble/droplet system, improving the attachment efficiency or altering the generated
bubble size.

However other test work has indicated an opposite effect, indicating a decrease in hydrocarbon removal
efficiency with salinity, though this behaviour is thought to be related to the saline water interfering with the
chemical flotation aids used.

Though the effect is variable, it is clear that salinity will have an influence on the flotation behaviour. This
is particularly important in systems which are exposed to large seasonal variations in water salinity due to
rainwater ingress. It would be expected that such systems would show a seasonal variation in flotation per-
formance.

Direct gas sparging


Test work, conducted by Shell Oil in 1975, indicated that higher hydrocarbon removal efficiencies than
traditional induced gas systems could be obtained by direct gas sparging.

Direct gas sparging has the advantage of giving the user control over the gas bubble concentration and gas
bubble size, allowing the flotation efficiency to be optimised. In addition, the sparging action is more gentle
than induced gas systems, improving the bubble/droplet attachment efficiency. The use of a lean sparging
gas may allow the removal of dissolved hydrocarbons from the water.

However, gas sparging has the disadvantage of requiring a suitable source of compressed gas for sparging,
possibly requiring a small compressor to either provide a source of gas or recover sparged hydrocarbon gas.

7.9.3. Installation/Configuration
It is recommended that gas flotation units are used as a second or third stage of a water treatment system. A
primary separation stage should be provided upstream, such as a plate interceptor, to provide for removal of
suspended solids and slugs of hydrocarbons.

An IGF flotation cell is effectively a well mixed vessel and thus a certain fraction of the feed stream will short
circuit through the cell. To achieve the required hydrocarbon removal efficiency a number of cells (typically
four) are combined in series to form a single package. A discharge box is provided after the last flotation cell
and is the usual location for water level control instrumentation. An input box can also be used to provide for
skimming of free oil and the removal of solids, though this is usually more efficiently performed in upstream
vessels. A typical four cell design is illustrated in Figure 7.9.2.

Although the majority of flotation packages operate at near atmospheric pressure, designs incorporated into
pressure vessels are available from some vendors. Pressurised designs may simplify some design aspects by
eliminating the need for pumps on outlet streams or the need to arrange the layout to achieve gravity flow.
Some vendors market smaller flotation packages consisting of one or two cell configurations. These could
be used for duties with relatively low inlet hydrocarbon concentrations, or could be used to add pre or post
flotation to an existing flotation installation. The MiSwaco (formerly Epcon) Compact Flotation Unit falls in
this category.

Deoiling Manual V 2.0 107


Restricted

The Epcon technology uses one or two flotation stages, performed in vertical vessels. The feed is entered via
a tangential inlet, establishing a mild rotation of the liquid in the vessel. Nitrogen gas is entered via a bottom
sparger. Assuming a separation efficiency per vessel of 60% an overall oil removal of the order of 80% can
be expected.

Vendors are usually responsible for the sizing and design of flotation packages. As each application and
associated water chemistry is different, field trials with trial units or bench scale systems are recommended to
ensure that flotation will be effective and to identify the potential need for flotation aids.

7.9.5. Operational considerations


Solids handling
As previously discussed, flotation cells should not be used for the removal of solids. The drainage system from
the cells should be suitable for the discharge of any solids or sludges that do accumulate, however, in general
the majority of solids should preferably be removed by upstream treatment equipment.

Chemical dosage rates


Experience has indicated that the overdosing of flotation aids can result in the formation of “jelly like” sub-
stances in the froth overflow. These can accumulate on overflow weirs, paddles and in launders. Care should
be taken to avoid overdosing of chemicals, particularly during turn-down. In many situations adequate flota-
tion performance can be achieved without the addition of chemicals.

Froth height
A characteristic of gas flotation is that the froth height is generally higher in the end cells with the lower hy-
drocarbon content. To some degree, froth height can be altered by changing the height of the froth overflow
weirs. Further reductions in froth height can be made through controlling the gas intake rate into the last cells,
either through partial plugging of the gas intake ports or the fitting of valves to allow the gas intake rate to
be manually adjusted as required.

Eductor erosion
The eductor systems for hydraulically induced flotation systems can be subject to blockage or erosion, espe-
cially if solids are present in the recirculation water. This can result in poor bubble formation and a deteriora-
tion in hydrocarbon removal performance. Eductors should be inspected during shutdowns to ensure there is
not excessive wear.

Operating experience
Induced gas flotation is a well established technology with numerous installations throughout the world.

108 Deoiling Manual V 2.0


Restricted

7.10. Dissolved gas flotation


7.10.1. Introduction
Gas flotation units are categorised by the method used to generate gas bubbles, either Induced Gas Flotation
(IGF) (mechanical and hydraulic) or Dissolved Gas Flotation (DGF). Dissolved gas flotation is discussed in this
section. Induced gas flotation is discussed in section 7.9.

As with induced gas flotation, dissolved gas flotation assists the removal of hydrocarbon droplets by attach-
ment to rising gas bubbles. The oily froth which forms on the surface of the water is removed by skimming
or overflowing to a collection trough. Gas flotation units are generally installed as a secondary stage in a
water treatment system.

Inert gases, air or hydrocarbon gas are normally used for induced gas flotation. Operation with air is often
referred to as Dissolved Air Flotation (DAF). In many cases the use of air is not practical due to the need to
exclude oxygen from the process (the presence of oxygen leads to concerns with corrosion, biological growth
and explosion hazard aspects). Chemical additives are often used to improve the performance of the flotation
process.

Figure 7.10.1 illustrates the design of one style of dissolved gas flotation system. Bubbles are generated by
saturating a liquid stream with gas, typically at a pressure of 3 to 6 barg. As the liquid is depressurised, the
gas is released as fine bubbles. The rising bubbles can attach themselves to hydrocarbon droplets and solid
particles and float them to the water surface from where they can be removed.

Figure 7.10.1: Dissolved gas flotation system.

The main advantages of the dissolved gas process over induced gas systems is the relatively gentle method
of generating the gas bubbles. The absence of high shear forces in the flotation cell allows DGF to handle
flocculated feed streams without breaking up the flocs.

For unassisted water treatment an induced gas flotation system is more efficient for removal of hydrocarbons
than a dissolved gas flotation system. However, the dissolved gas system has been found to be more efficient
in removing fine solids. As a result, a dissolved gas flotation system would often be recommended over an

Deoiling Manual V 2.0 109


Restricted

induced gas system for applications requiring solids removal or where a pre-flotation flocculation stage is
used or required.

Figure 7.10.1 illustrates the use of a stirred vessel as the flocculation stage for the IGF system. Upstream of the
actual flotation process the feed passes through a flocculation tank. To enhance the separation efficiency the
droplet size is increased by means of an agglomeration step. Suspended solids and hydrocarbon droplets
in the feed stream are pulled together to form larger agglomerates called microflocs. This process is assisted
through the addition of a coagulant, typically ferric chloride, aluminum sulphate or ferric sulphate. In the
subsequent flocculation step the microflocs formed by coagulation are further agglomerated into larger flocs.
This process is assisted by residence time, gentle agitation and the addition of polyelectrolytes (flocculants).

Chemicals can also be used to assist the removal of solids, by helping to bind the solids into
the flocs, or by altering the surface chemistry to make the solids more hydrophilic and thus more likely to
be gathered by the flotation bubbles.

Figure 7.10.2 illustrates a common variation of dissolved gas flotation system, the use of a plate pack to
improve the hydrocarbon separation performance.

Figure 7.10.2: Dissolved gas flotation unit with plate interceptor and serpentine pipe flocculator.

The plate pack provides a laminar flow regime which allows small hydrocarbon droplets, flocs and gas
bubbles to be separated from the water stream. Also illustrated is a serpentine pipe flocculator. This type of
flocculator operates in plug flow, using the natural flow turbulence and the residence time in the serpentine
pipe system to promote flocculation.

The stream to be saturated with gas can be the full feed stream, a proportion of the feed stream or a recycle
stream from the discharge of the DGF unit. The recycle stream system is the most common, having the advan-
tage of using a clean water stream in the saturation system (i.e. no solids or hydrocarbons to contaminate
the saturation equipment) and ensures that the largely hydrocarbon free recycle stream, not the feed stream,
is exposed to the shear forces experienced during the depressurisation of the gas saturated stream. This is
particularly important in ensuring the maximum hydrocarbon droplet or floc size enters the flotation cell.

110 Deoiling Manual V 2.0


Restricted

7.10.2. Performance variables


Hydrocarbon droplet size
It is recommended that the DGF process be used in combination with an upstream flocculation system. This
will gather small hydrocarbon droplets into large flocs, improving the separation efficiency. By incorporating
the flocculation stage the DGF system should be relatively independent of the hydrocarbon droplet size in
the feed stream.

Field trials have indicated that a DGF unit can achieve a hydrocarbon removal efficiency in the order of 90%
or higher with a suitable flocculation stage. The use of flocculation may also assist the removal of fine solids
from the water stream.

If a flocculation stage is not used, an induced gas flotation system as described in section 7.9 would gener-
ally be recommended over a DGF unit.

Bubble size and concentration


The performance of gas flotation is improved by reducing the size of the bubbles and increasing the number
of bubbles (bubble concentration). The typical bubble size for a DGF unit is in the order of 40 to 70 microns.
This is smaller than the typical 100 to 400 mm bubble size in an induced gas flotation system. To minimise
bubble size it is important to avoid any premature release of gas in the piping between the saturator and
the flotation cell, and to distribute the bubbles widely into the flotation cell. This minimises bubble size by
avoiding bubble contact and coalescence. Excessive bubble coalescence or a high bubble concentration
will result in the formation of slugs of gas, with increased turbulence in the cell and reduced separation per-
formance.

The solubility of the gas in the water will depend on a number of factors such as the temperature of the water,
the gas being used for saturation and the influence of impurities in the water (hydrocarbons, electrolytes, etc.).
Figure 7.10.3 illustrates the equilibrium solubility of a selection of gases. It can be seen that the hydrocarbon
gases are more soluble than air, and that the solubility of the gases decreases with increasing temperature.

Figure 7.10.3: Solubility of gases in water.

Deoiling Manual V 2.0 111


Restricted

The amount of gas released during depressurisation will be a function of the difference in gas solubility
between the pressurised and depressurised water and the amount of gas that can actually dissolve into the
water in the saturator. Due to the short contact time, equilibrium gas solubility will not be achieved between
gas and liquid in the saturator. A saturation efficiency in the order of 70% can be expected for a simple
saturation vessel, increasing to the order of 90% for a more efficient saturator using internal packings or a
pump based saturator.

Inlet oil concentration


It is recommended that the inlet oil concentration to a DGF unit is kept lower than 500 mg/l. It is expected
that as for induced gas flotation systems, very high concentrations of hydrocarbons or slugs of pure hydrocar-
bons would upset the performance of the DGF unit.

As for induced gas flotation systems, a primary treatment stage will often be used upstream of the DGF sys-
tem. The primary treatment stage should smooth the flow to the DGF and remove large hydrocarbon droplets,
hydrocarbon slugs and large solid particles.

As with most deoiling equipment, the hydrocarbon concentration of the effluent water will be related to the
inlet hydrocarbon concentration by the hydrocarbon removal efficiency. Even with high removal efficiencies,
a high inlet concentration will lead to a high outlet concentration.

Recycle rate
The recycle rate of clean water for pressurisation is typically in the order of 10-30% of the feed rate. Increas-
ing the recycle rate will increase the bubble concentration in the flotation cell, however at the expense of
residence time in the flotation cell.

Flow rate
Altering the flow rate through the flotation cell will alter the separation performance by changing the resi-
dence time. The longer the residence time, the better the performance.

It should be noted that most of the floc/bubble contact will often take place in the inlet piping or inlet cham-
ber of the DGF where the bubble concentration is highest. The residence time in the inlet system is often very
short, in the order of 20-30 seconds. Outside of the inlet system the bubble concentration and chances of
droplet/bubble contact fall rapidly. Thus most of the hydrocarbon/solids removal is accomplished in the inlet
system with the bulk of the residence time in the DGF allowing for rising of the flocs and bubbles.

Flotation aids
Trial work has indicated that in some cases the performance of gas flotation may be considerably enhanced
by using a suitable flotation aid.

Flotation aids may act through several mechanisms such as promoting the coalescence of the dispersed hy-
drocarbon droplets, enhancing the attraction between the hydrocarbon droplets and the rising gas bubbles
or improving bubble and froth stability.

Flotation aids should be injected well upstream of the flotation cells to ensure effective mixing and to provide
time for the chemical to act. General experience with flotation aids indicate that there is an optimum dosage
rate. At low dosage rates performance is poor, but improves as the injection rate increases. At excessively
high dosage rates the performance falls off again and the chemical can react with the froth to form “jelly like”
compounds that can accumulate in the froth discharge area.

The correct flotation aid and dosage rate can only be selected on the basis of field trials. Typical dosage
rates are in the order of 5 to 30 mg/l.

112 Deoiling Manual V 2.0


Restricted

Process chemicals
The effect of oil field chemicals on flotation performance can only be effectively investigated through field
trials. However experience indicates that the presence of corrosion inhibitor can dramatically reduce the
performance of a flotation unit. The mechanism for this interference is not clear, but could be either the stabi-
lisation of very small hydrocarbon droplet sizes or interference with the attraction between gas bubbles and
hydrocarbon droplets.

Suspended solids
The effect of suspended solids on the flotation performance should be carefully considered. In some cases
solids can be effectively floated out of the cells with the froth, allowing the flotation cell to remove both solids
and hydrocarbons.

However in other cases poor performance of flotation systems has been experienced with water containing
high levels of suspended solids. In these cases the flotation action will carry some solids into the froth, but will
also prevent other solids settling and these will then be carried out with the water stream. Solids may carry
hydrocarbon droplets or hydrocarbon films, contaminating the effluent water.

It is recommended that the upstream primary treatment stage is designed for the removal of the larger, rapid
settling solid particles. With the correct addition of flotation chemical aids, the remaining fine solids should
be successfully removed by the flotation process.

7.10.3. Installation/configuration
The flotation process is typically a secondary treatment process. A primary separation stage should be pro-
vided upstream, such as a plate interceptor, to reduce the inlet oil concentration, and to provide for removal
of suspended solids and slugs of hydrocarbons.

The DGF process requires a liquid stream to be saturated with gas under pressure which will then provide a
source of flotation bubbles when the pressure is released. A number of different pressurisation schemes are
illustrated in Figures 7.10.4 to 7.10.6.

Full flow operation


Full flow operation, illustrated in Figure 7.10.4, consists of pressurising and dissolving the flotation gas in the
entire feed stream. This has the advantage of generating the maximum quantity of bubbles in the flotation cell.

This system has the disadvantage that the entire feed stream must be depressurised across a valve to gener-
ate the required gas bubbles afler pressurisation. De-
pending on the feed stream pressure, a pump may
also be required to pressurise the feed stream.

These pressure changes make hydrocarbon separa-


tion more difficult due to shearing of hydrocarbon
droplets or the break up of flocs.

Split flow operation


Split flow operation, illustrated in Figure 7.10.5, Figure 7.10.4: Full flow operation.
only pressurises part of the feed stream. This has the
advantage of a smaller pressurisation pump and al-
lows the use of a flocculation stage which will improve the flotation performance.

However split flow has the disadvantage of generating less bubbles in the flotation cell and still shearing the
hydrocarbon droplets and flocs in the depressurisation of the split feed stream.

Deoiling Manual V 2.0 113


Restricted

Recycle operation
Recycle operation, illustrated in Figure 7.10.6, is the
most common mode of operation. Recycle operation
has the advantage of minimising upstream shearing
of any of the feed stream and allowing optimum floc-
culation.

It also has the advantage of allowing more sophisti-


cated saturator designs as the recycle water is rela-
tively free from solids and hydrocarbons.

However it has the disadvantage of a lower bubble


Figure 7.10.5: Split flow operation.
concentration within the flotation cell as the recycle
stream is only 10-30% of the feed stream. Due to the
recycle, a larger flotation cell will be required for the
same liquid flow rate.

Feed stream from upstream process


One potential source of a gas saturated liquid stream
is the process itself. This technique was applied on
the North Sea Beatrice field in the conversion of a
conventional induced gas flotation package to dis-
solved gas flotation. The conventional bubble induc-
ers were stripped from the package and gas bub-
bles were provided by a controlled depressurisation Figure 7.10.6: Recycle stream operation.
st
of production water from the 1 stage separator.
Claimed benefits are increased throughput and 80%
hydrocarbon removal efficiency. A similar effect oc-
curs in degassing vessels. The degassing vessels at Draugen and Shearwater have been retrtofitted with
dedicated internals to make maximum use of this effect, see par 6.4.

Intermittent skimming
The API guide to water treatment claims one operating installation found an improvement in DGF perform-
ance by operating with intermittent skimming. This was believed to allow the formation of a stiffer, relatively
stable and more concentrated foam/scum layer on the water rather than damaging the foam/scum layer
with continuous skimming.

114 Deoiling Manual V 2.0


Restricted

7.11. Deep bed media filtration


7.11.1. Introduction
Deep bed media filtration involves the removal of suspended hydrocarbons and solids from waste water by
passing the water through a deep bed of granular material, resulting in a clean product stream (the filtrate).
Media filters are used as polishing water treatment technology for deep removal of oil and/or suspended
solids. These technologies typically require upstream secondary treatment, such that the oil and solids level
in the feed is less than 50-100ppm.

As with most deoiling equipment, the hydrocarbon removal efficiency of the filter will be a function of the hy-
drocarbon droplet size in the feed. However for most applications media filters can produce a water effluent
with a hydrocarbon content of less than 10 mg/l dispersed hydrocarbons, often achieving less than 5 mg/l.

It is important to note that media beds do not operate by surface filtration. Surface filtration is the separation
mechanism where the particles or droplets to be filtered from the water removed become wedged in the
spaces between the grains of media. Surface filtration in a deep bed filter is undesirable as it will quickly
result in blockage of the top layer of the bed and the capacity of the subsequent layers of the media bed will
not be effectively utilised.

The pore openings between the granular media grains are relatively large compared to the suspended par-
ticles, allowing the particles to pass into the bed of the filter. Inside the bed the suspended particles become
trapped by first intercepting, then attaching to the media. Thus the suspended particles and droplets in the
feed stream are removed within the body of the media bed, not on the surface.

Sand filters
The simplest form of media filter is the sand filter. The disadvantage of a sand filter is that when the filter is
cleaned by backflushing, the sand particles will tend to segregate with the smallest sand particles concen-
trated in the uppermost layers, Whilst this segregation is ideal for a filter in which the water flows upwards,
in the majority of media filters the water flow is downwards. With down flow operation this surface layer of
fine sand will block more rapidly than larger sand particles, reducing the length of the filtration cycle.

This segregation of the sand particles during the backwash operation is partly alleviated through the use of
multimedia filters.

Multimedia filters
A typical multimedia sand filter design is il-
lustrated in Figure 7.11.1.

In a multimedia filter, two or more media with


a different grain size and density are used
with the denser particles having the smaller
grain size. During backwashing, the two
media will segregate with the larger sized,
less dense media forming a layer above the
dense, small sized media.
Within each media layer there will still be
size segregation, however the overall per-
formance of the filter will be improved. The
larger particles on the surface of the filter will
minimise blockage of the surface layers, al-
lowing more of the bed depth to be utilised Figure 7.11.1: Multimedia down flow sand filter.

for filtration.

Deoiling Manual V 2.0 115


Restricted

Nut shell media filters


An alternative filter media is crushed nut shells, such as the Hydromation filter (walnut shells) or the Wemco
Silver Band filter (walnut/pecan shells). This media is preferentially wetted by the water phase which makes
the media easier to clean during the backwash cycle and improves the long term performance of the filters
when used for the filtration of hydrocarbons.

Because of their potentially superior performance in deoiling waste water, nut shell media filters are recom-
mended over conventional media filters for the removal of dispersed hydrocarbons from water. These filters
are discussed in more detail in section 7.11.6.

Absorption media filters


Several companies (Symons, Twinfilter, Cetco) offer produced water treating packages with deep bed media
absorption filters, specifically for the removal of dispersed oil and a range of dissolved hydrocarbons from
produced water of gas production platforms. The media applied are based on organic or organoclay sub-
strates such as cellulose, bentonite, anthracite and palygorskite, impregnated with proprietary surfactants.
PS-85, based on palygorskite, appears to perform the best in terms of handleability. This is also used in the
Twinfilter Oilclog system where it is contained in cartridges instead of in the form of a bed in a vessel.
The Oilclog system can reduce the dispersed oil concentration to below 10 mg/l provided that the level in
the feed does not exceed 100 mg/l.
The absorption media filters are discussed in more detail in section 7.11.7.

Cleaning of media filters


Over the length of the filtration cycle, the media bed of the filter will become progressively blocked by ac-
cumulated solids and hydrocarbons. Eventually the media bed will require cleaning, either when the flow rate
through the filter falls to an unacceptable level, when the pressure drop across the filter becomes excessive
or when the quality of the filtrate deteriorates.

Cleaning of media filters is generally conducted by taking the filter off-line and fluidising the media bed with
a back flow of clean water, allowing the accumulated hydrocarbons and particles to be washed away. The
cleaning of the media is enhanced by collisions between the fluidised media and is often assisted by sparg-
ing air or gas into the filter to scour the media. Hot water backflushing and/or surfactant chemicals may be
used to assist the removal of hydrocarbons from the media.

Normally clean, filtered water is used for backwashing and provision may have to be made to store a suit-
able quantity of backwash water. Alternatively clean water may be taken from the discharge of another on-
line fitter, however this will require the filter package to effectively have two spare filters, one being cleaned
and one providing water for backflushing.

The crushed nut shell media filters do not require clean water for backwashing and their special cleaning
arrangements are discussed in more detail in section 7.11.6.

7.11.2. Performance variables


Hydrocarbon droplet size and solid particle size
In general terms, the larger the size of a hydrocarbon droplet or solid particle, the greater the probability
that it will be removed by a deep bed media filter. This is illustrated in Figure 7.11.2 where it can be seen
that the separation efficiency rises with particle size. Media filters are typically capable of removal of 95%
of particles 2 µm or larger and 98% of particles 5 µm or larger.

However, the performance of a filter must be considered in relation to both particle removal efficiency and
overall filtration efficiency. Particle removal efficiency is a measure of how efficiently a filter removes a particle
or droplet, regardless of the throughput through the filter. Filtration efficiency takes the flow rate through the
filter into consideration.

116 Deoiling Manual V 2.0


Restricted

Figure 7.11.2: Typical separation performance of a multimedia filter. .

For example, large solid particles or droplets entering a media filter will be quickly removed in the uppermost
layer of the media bed, leading to surface filtration. The surface layer will be very efficient at removing incom-
ing particles, hence will have a high particle removal efficiency. However, the actual flow rate through the
surface layer will be restricted, resulting in a low overall filtration efficiency.

This behaviour is illustrated in Figure


7.11.3. It can be seen that above
a certain ratio of particle diameter
to media diameter, surface filtration
effects will dominate and although
the particle or droplet removal ef-
ficiency will continue to rise, the
overall filtration efficiency will fall.

Thus, for a given media grain size,


there will be an optimum feed par-
ticle or droplet size that will be ef-
ficiently filtered without inducing
surface filtration. As an indication,
for a media bed of spherical par-
ticles, the largest particle that can
pass into the bed has a diameter
of about one seventh (1⁄7) of the di-
ameter of the media particles. Thus
the feed to a filter consisting of 0.5
Figure 7.11.3: Particle removal efficiency and filtration efficiency versus ratio of feed particle or
mm sand particles should not con-
droplet size to filter media size. .
tain particles or droplets larger than
70 microns. In general it is recom-

Deoiling Manual V 2.0 117


Restricted

mended that particles or droplets larger than 100 µm are not allowed to enter the filter. This particle size
corresponds to the size of solid particle or oil droplet that is readily removed by relatively simple upstream
equipment such as a plate pack interceptor.

Flow rate
A filter should not be operated above its recommended maximum flow rate. Operating a filter at capacities
above the design flow rate will raise the velocity though the open pores in the media bed. This will tend to
detach the particles and hydrocarbons captured in the media bed and carry them through the bed into the
filtrate, resulting in a deterioration of the filtrate quality and a shortening of the length of the filtration cycle.

Pressure drop
The pressure drop across a clean media filter is typically in the order of 0.4 bar. When used for filtering solid
particles the pressure increase over the filtration cycle is normally in the order of 0.3 to 1.4 bar. When the
filtration cycle is dominated by hydrocarbon removal, the pressure increase is generally at the low end of this
range when hydrocarbon breakthrough into the filtrate occurs.

Filtration cycle
Filtration cycle consists of 5 stages:
■■ Filtration
■■ Media agitation
■■ Backwash
■■ Pause/ sedimentation
■■ Pre-run flush

The length of the filtration cycle is determined by either an unacceptable rise in differential pressure across
the filter as the media bed becomes blocked or through excessive levels of contamination in the filtrate. For
the filtration of solids, the rise in the pressure differential across the filter is likely to be limiting, while for the
filtration of dispersed hydrocarbons the quality of the filtrate is likely to be limiting.

It should be noted that when differential pressure is the limiting factor, the length of the filtration cycle will
decrease dramatically with increasing flow rate as the increasing flow rate has a double impact. Firstly, forc-
ing a higher flow rate through the filter will itself result in a higher pressure drop (ΔP~v2). Secondly, the higher
flow rate will mean that the rate of solids entering the filter will be higher, blocking the filter more rapidly.

Thus a doubling of the flux rate through the filter can lead to a four fold reduction in the length of the filtration
cycle for a differential pressure limited filter.

Inlet hydrocarbon and solids concentration


Media filters are not tolerant of high concentrations of hydrocarbons or solids in the feed. High concentra-
tions of hydrocarbons in the feed will tend to coalesce into larger droplets, increasing the probability that the
hydrocarbons will be trapped in the surface layers of the bed, rapidly blocking the filter.

The acceptable inlet hydrocarbon content is a compromise between the length of the filtration cycle and the
nature of the hydrocarbon. Multimedia filters are claimed to be suitable for feed hydrocarbon concentrations
in the order of 100 mg/l, however it is recommended that the inlet hydrocarbon content is normally limited
to the order of 30-50 mg/l. Nut shell media filters are suitable for hydrocarbon concentrations in the order
of 100-150 mg/l as they are more easily cleaned.

High solids concentrations have a similar surface filtration effect due to bridging. A number of solid particles
can wedge together, effectively bridging over and blocking the pores in the media bed. The solids concentra-
tion in the feed should typically be less than 50 mg/l.

118 Deoiling Manual V 2.0


Restricted

Hydrocarbon characteristics
The physical characteristics of the hydrocarbon will have a significant effect on the length of the filtration
cycle. Heavy oils are more likely to block the surface layers of the filter, shortening the filtration cycle.

Similarly, waxy or fouling hydrocarbons will reduce the length of the filtration cycle and will shorten the
expected useful life of the filter media. In some cases, filters can become clogged by oil accumulation and
cannot be cleaned by backwashing. In other cases the filter media can be aggglomerated by the fouling,
collecting the media into “mudballs”.

Vent stream
Depending on the filter configuration, larger hydrocarbon droplets or gas bubbles will tend to rise and collect
in the top of the filter vessel, This accumulation is removed by maintaining a small flow from a vent at the top
of the filter vessel. The flow rate of this vent stream is usually manually adjusted on the basis of operational
experience. A suitable flow path must be identified for the recycling of any vent stream back to the process.

Treatment of backwash water


During backwashing a relatively large reject stream is generated containing the filtered hydrocarbons and
solids. The backwash stream cannot be recycled directly to the process as the solids would lead to solids
accumulation in the process system, and to emulsion problems. Instead the backwash stream is generally
directed to a clarification vessel which allows the solids to settle and the hydrocarbons to rise. Separated
hydrocarbons and accumulated solids/sludge are periodically removed from the clarification vessel. The
clarified water can be recycled to the process.

The volume required in the clarification tank may be a undesirable weight penalty in an offshore installation.
The volume of backwash water requiring clarification could potentially be reduced by passing the backwash
stream though a solid/liquid hydrocyclone to remove the solids. The solids/water stream from the cyclone
underflow could be directed to a smaller clarifier while the hydrocarbon/water overflow stream could be
recycled to the process. Alternatively, consideration should be given to the crushed nut media filters which
have significantly lower backwashing requirements.

7.11.3. Installation/Configuration
Process location
As media filters are not tolerant of high hydrocarbon or suspended solid concentrations in the feed stream
they are generally used as the third stage of water treatment. As previously discussed, it is recommended
that upstream equipment should be designed to remove hydrocarbon droplets and solid particles larger than
100 µm. Course strainers (100 µm) should also be installed on the backwash system to prevent blockage of
the filter distribution and collection systems.

Packaging
Media filters are typically packaged into vertical pressure vessels. Suitable distribution and collection inter-
nals are required to ensure an even distribution of the water flow though the media bed.

One standby filter should be provided to be brought on-line when other filters are backwashed or shut- down
for maintenance.

When specifying a filtration system, it should be remembered that in addition to the filter vessels themselves,
provision may be required for a clarification tank for the backwash water and a storage tank to store clean
filtered water to perform the backwash.

Size, Weight and Capacity


The typical operating weight for a skid mounted packaged deep bed filtration system for the treatment of
oily water is 25 ton per 100 m3/h capacity. The figure may be used for both dual media filter systems and

Deoiling Manual V 2.0 119


Restricted

crushed nut shell filter designs. However, additional weight allowance should be added as required for aux-
iliary systems such as chemical injection facilities, backwash water storage and clarification tanks.

7.11.4. Design guidelines


The following guidelines can be used for indicative sizing of downflow multimedia filters for the filtration of
oily water. Higher flux rates are possible for sea water filtration duties, see the WATER FLOOD MANUAL.
Guidelines for crushed shell filter media designs are given in section 7.11.6.
Flux rate through filter: 20-40 (m3/h)/m2
Accumulated hydrocarbon capacity of filter bed before cleaning: 40 kg/m2
Accumulated solids capacity of filter bed before cleaning: 20 kg/m2
Maximum recommended continuous hydrocarbon feed concentration: 30 -50 mg/l
Multimedia backwashing water flow: 40 (m3/h)/m2 for 15 to 30 minutes
Multimedia Backwash water volume: Typically 3% of feed
The rate of media loss from the filter is a function of the frequency of backwashing and proper control of the
backwash cycle. Media is also lost by attrition of the media particles. Typical media loss is in the order of
10% per year.

When filtering oily water, the dirty backwash water is generally recycled to the process after the hydrocar-
bons and solids are removed. The capacity of the filtration package has to be sufficient to accommodate this
recycling flow.

The frequency of backwashing should be such that there is sufficient time in the clarification tank to allow sol-
ids and hydrocarbons to settle from the backwash water and for clarified water to be pumped away before
the next filter backwash.

Backwashing control
During the backwash cycle the media bed is fluidised. The degree of fluidisation is a function of the density
and the viscosity of the backwash fluid, both of these variables being functions of the backflush fluid tem-
perature.
In some designs air scour is used in the backwash cycle to agitate the media bed, and separate the dirt from
the media grains.

The optimum backwash flow rate to give sufficient fluidisation without excessive media loss should be de-
termined for the range of backwash fluid conditions, The backwash fluid flow rate should be measured and
controlled at the appropriate rate. If backwash fluid conditions are likely to change significantly (e.g. wide
temperature variations) then automatic adjustment of the backwash flow rate setpoint may be warranted.

7.11.5. Operating experience


A large number of deep bed media filters are commonly used throughout the petroleum industry for the filtra-
tion of sea water for water injection or for produced water clean-up. However, there are only limited applica-
tion installed on offshore facilities.

7.11.6. Nut shell deep bed media filters


7.11.6.1. General
The nut shell filtration was developed as a more suitable method of filtering free oil in applications where
sand filters have traditionally been used. Nut shell filters are a form of deep bed media filter, utilising crushed
walnut or pecan nut shells as the filter media. The Walnut shell media have an equal affinity for oil and water.
As the Oil is adsorbed not absorbed, the media can be recharged by means of backwashing. This ability to
deal effectively with hydrocarbons in the feed makes the nut shell filters more suitable to deoiling duties than
other deep bed media filters.
.

120 Deoiling Manual V 2.0


Restricted

Backwashing is performed using the unfiltered feed water as the fluidising medium, eliminating the require-
ment for a storage tank to store filtered water for backwashing. Depending on the design, gas or air is used
in the backwashing as well. Sufficient hold-up volume in the upstream system (& downstream system if the
water is to be used for PWRI) is required to cope with the filter backwash cycle. The flow rate of backwash
water is significantly lower than for conventional deep bed filter designs, minimising the size of additional
facilities required to deal with the dirty backwash water.

As the media is efficiently cleaned there should be no requirement for media replacement except to cover a
10% annual media loss due to attrition.

7.11.6.2. Hydromation®
The Hydromation® filter is illustrated in Figure 7.11.4. Hydromation® is a registered trademark of Cameron
International Corporation. Hydromation® filters are down flow, deep bed, walnut shell media filters,designed
for produced water filtration.

During the backwash cycle the filter is shut in and the media scrubber pump used to fluidise, then circulate
the filter media downwards through the external media scrubber. The media scrubber contains a wedge-wire
strainer that allows small particles and hydrocarbons to pass to the backwash discharge, while the filter me-
dia is returned to the filter vessel. The turbulence and collisions generated during the backwash recirculation
clean the filter media.

After the backwash the flow through the filter is reversed and the filter bed reforms. The filter is then operated
for a short period on recycle to wash the contaminated backwash water out of the filter bed. The filter is then
clean and ready to be brought back on line.

Figure 7.11.4: Hydromation® crushed shell deep bed media filter.

Deoiling Manual V 2.0 121


Restricted

The Hydromation® filter is available in two configurations. The down flow configuration illustrated in Figure
7.11.4 for capacities up to 350 m3/h and a radial flow configuration for capacities up to 1,025 m3/h. The
radial flow configuration is more space and weight efficient, but is only suitable for relatively high flow rates.

7.11.6.3. Wemco Silver Band


The Wemco Silver Band filter is illustrated in Figure 7.11.5. During the filtration cycle the filter acts as a nor-
mal deep bed filter, using a mixture of crushed walnut and pecan nut shells as the filter media.
The backwashing and media cleaning system is similar to the Hydromation filter with the clean water outlet
valve shut and the filter media recirculated by a pump past a separator screen. The screen allows the solids
and hydrocarbons to exit the filter but retains the filter media inside the vessel. The Silver Band filter has the
separator screen built inside the filter vessel itself.

Figure 7.11.5: Wemco Silver Band crushed shell deep bed media filter.

After the backwash the recirculation is stopped and the filter bed reformed. The filter is then clean and ready
to be brought back on line.

The Wemco Silver Band filter has been tested at the Orkney Water Test Centre. The filter was found to consist-
ently achieve in the order of 10 mg/l hydrocarbon in the effluent from a 200 mg/l feed stream.

7.11.6.4 Walnut shell filter SIemens


Siemens Water Technologies provides the Auto Walnut Shell (AWS) Filter as well, using a deep bed of black
walnut shells. Recently, Siemens Water Technology has introduced the more compact pulsed bed walnut shell
filter which is specifically intended for offshore application. The backwash flow is minimized by pulsing the

122 Deoiling Manual V 2.0


Restricted

flow and by mixing it with gas. At this moment the new system has only been pilot tested at BP Valhal; there
are no applications installed on off-shore facilities.

7.11.6.5. Performance characteristics


A walnut shell filter can remove dispersed hydrocarbons from up to 100 mg/l to below 10 mg/l. It will not
remove dissolved hydrocarbons such as organic acids, phenols and aromatics. When the oil droplets are
small, coagulants can be used to increase the effective droplet size and improve separation.

Solids will be removed to some extent as well, especially when coated with oil. The solids concentration
should be limited to < 200 mg/l to prevent rapid pressure build up. Solids removal efficiencies claimed vary
from 65% (Hydromation) to 90% (Siemens). To be more precise:
solids removal
TSS (mg/l)
efficiency (%)
Hydromation < 200 65

< 20 95 reduced flow rate

Siemens < 50 90

During the actual filtration cycle, the response of a nut shell filter is not different from other media filters. High
hydrocarbon concentrations result in surface filtration and shortening of the filtration cycle, especially with
heavy crude oils. However, due to the hydrophilic nature of the filter media, the crushed shell filters are more
efficiently cleaned and return to good performance after the regeneration cycle. This ability to deal effectively
with hydrocarbons in the feed makes the walnut shell filters more suitable to deoiling duties than other deep
bed media filters.

Problems have been experienced when ferric suphide salts precipitate in the filter bed due to the mixing of dif-
ferent water sources. The precipitate attaches to the walnut particles and reduces their ability to regenerate.
If the bed is contaminated in this way the salts can be removed by washing with acid. Further precipitation
can be prevented by adding small quantities of acid.

7.11.6.6 Design guidelines


The following guidelines can be used for indicative sizing of crushed shell media filters.
■■ Flux rate through filter: 25 -40 (m3/h)/m2
■■ Maximum recommended continuous hydrocarbon feed concentration: 50-100 mg/l
■■ Total backwash cycle time- 15-20 minutes
■■ Backwash water volume- Typically less than 1% of throughput (Hydromation®: 0.5-1%; Siemens: 0.4-0.7%)
■■ Backwash frequency: 1 and 2 backwashes per day (minimum is once per day)
■■ Operating temperature should be less than 80 °C
■■ Specific gravity of nut shell media: 1.3-1.4

Sufficient buffer volume must be available in the upstream system to handle the backwash flow. The backwash
handling system should be designed to cater with at least 2 backwashes. If the water is to be re-injected,
buffer volume is required downstream as well.

Media is lost by attrition of the media particles. Typical media loss is in the order of 5-10% per year. For the
Hydromation Deep Bed Nutshell Filter a maximum of 5% attrition of media per annum is guaranteed, but
typical media loss is less than 1 to 2% per annum.

Deoiling Manual V 2.0 123


Restricted

7.11.6.7. Size and weight comparison


The size and weight of a crushed shell filter system is approximately the same as an equivalent dual media
filtration system.

The crushed nut shell filters do offer some external weight savings in that a clean backwash water storage
tank is not required and that due to the smaller quantity of backwash water generated, the backwash water
clarification tank does not need to be as large as required for an equivalent dual media system.

7.11.6.8. Application
Walnut shell filters are widely applied onshore. This due to the size and weight of the filter and the more
stringent on-shore disposal limits. At Drilling Station Mittelplatte offshore Germany, RWE DEA operate a wal-
nut filter downstream of hydrocyclones to treat produced water for reinjection, which reduces the oil level to
below 5 mg/l.

7.11.7. Absorption media filters


Absorption media are either based on bentonite (or a blend of bentonite) or Palygorskite (also known as
Attapulgite). The media material is impregnated with surfactants. The type of surfactant will determine the
removal efficiency of the hydrocarbons. The absorption media filters are often referred to as organoclays.
Organoclays are mostly used for dispersed hydrocarbon removal, when oil level in feed is less than 100 mg/l.
Outlet concentration of less than 10 mg/l can be obtained. Once the media is spent, it needs to be re-
placed. Large OPEX costs are involved with significant dispersed oil levels (> 100mg/l). Some dissolved
hydrocarbons can be removed as well, but data are limited.

One major difference is that Palygorskite is a non-swelling clay, whereas Bentonite clays all swell markedly
when wetted. The change-out of PS-85 media (based on Palygorskite) is less labour intensive compared to
RM-25, as the spent PS-85 media retains its loose characteristics.

The main suppliers are:


- Enhydra Ltd.: PS-85 (bulkmaterial, based on Palygorskite);
- RM Environmental: RM-25 (bulkmaterial, based on bentonite),;
- CETCO Oilfield services: Crudesorb Media (based on bentonite);
- Twinfilter (absorbent cartridges filled with PS-85)

7.11.7.1 Performance characteristics


Oil & Grease level below 10 mg/l can be obtained, when average O&G level in the feed is 100 mg/l
or less. Some removal of dissolved hydrocarbon can be expected, such as BTEX. Limited operational data
is available on dissolved hydrocarbon removal. Performance will depend on the mixture of hydrocarbons.
Phenols are difficult to remove by means of an absorbing media. Other hydrocarbons, such as aromatics
(BTEX), will impact the phenols removal efficiency as well, due to higher preference to the absorbing media.
Therefore it is hard to predict the removal efficiency in advance; the only way to know for sure is to test it in
the field.

Solids should be as low as possible to prevent pressure build-up. Cartridge filters can be installed upstream
adsorption unit. Acceptable solid content in the feed of cartridge filters is typical 5mg/l or lower.
Vendors claim that a high methanol level (above 2-3vol%) will remove the surfactant from the absorption me-
dia. The surfactant is required to extract the hydrocarbons from the water phase. Therefore, at high methanol
levels the removal efficiency of the absorption media will be lost rapidly.

124 Deoiling Manual V 2.0


Restricted

7.11.7.2. Applications
Absorption filters are installed at Malampaya (Shell Philippines), downstream the methanol recovery column.
SPEX tested the main absorbents for oil removal, i.e.
■■ RM-25 (based on bentonite);
■■ RM27P (based on bentonite, blended with 20-30% anthracite);
■■ Crudesorb (based on bentonite);
■■ PS-85 (based on Palygorskite).
At this stage Malampaya will stay with PS-85

Twinfilter’s OilClog system uses cartridges filled with PS-85. 50 cartridges are fitted in a vessel. This design
simplifies the change-out of the absorbent (change-out of 50 cartridges takes less than one hour). Typical
change-out frequency is 2-3 weeks. It is typically used to treat relatively small flows, i.e. of the order of
1 m3/h.

The OilClog system is applied by several operators at the North Sea (Total, Gaz de France, Wintershall,
VPN). Oilclog filters are installed at Sakhalin OPF as temperorary solution to treat 40 m3/hr produced water.

7.11.7.3. Design guidelines


■■ Hydrocarbon Holding Capacity: Can absorb 30% - 100% of its weight, depending upon hydrocarbon
concentrations and solubilities, and media type
■■ Specific Gravity of Media: 2.0 – 2.2
■■ Maximum Operating Temperature: 60-80 °C, depending on media type.
■■ Safe pH range is 4 to 11

Deoiling Manual V 2.0 125


Restricted

7.12. Membranes
Membranes act fundamentally like filters. Membrane processes cover microfiltration, ultrafiltration and nanofil-
tration. Microfiltration and ultrafiltration use porous membranes and separate on the basis of size, with drop-
lets or particles larger than the pore size not passing through the membrane. Nanofiltration uses non-porous
membranes and separates on the basis of selective solution and diffusion of a species through the membrane

Figure 7.12.1 illustrates the typical operating ranges of the various membrane filtration processes.
To separate hydrocarbons from water, membranes are operated in a cross flow configuration with the oily
water flowing across the membrane surface. The cross flow configuration helps minimise fouling by sweeping
hydrocarbons and solids along the membrane surface. The water at the membrane surface will preferentially
pass through the membrane, producing a clean water permeate. The remaining concentrated hydrocarbon/
water stream is known as the retentate or concentrate.

Figure 7.12.1: Range of membrane filtration processes.

As illustrated in Figure 7.12.2, membrane systems are normally operated with a recycle stream which can
be up to 20 times the feed rate. This raises the hydro-
carbon concentration of the stream feed across the
membrane surface and reduces the flow rate of the
reject stream that is withdrawn to remove the accumu-
lating hydrocarbons.

While membranes for oily water clean-up show con-


siderable promise, the generally rapid decline in flux
rates due to fouling and variable results in membrane
cleaning have hampered their application.

Recently positive leads have been reported for ce- Figure 7.12.2: Typical membrane filtration flow scheme.
ramic membranes, in particular for the separation of
very fine oil in water dispersions stabilized by high
corrosion inhibitor concentrations. Further exploration
is in progress.

126 Deoiling Manual V 2.0


Restricted

7.13 References
1. Ten Bosch, B.I.M., and K. Al Bataineh, ‘De-bottlenecking study of Yibal Produced Water Skim Tank Using
CFD’, OG.04.20125.

Deoiling Manual V 2.0 127


Restricted

Appendix 7.1

API Separators

128 Deoiling Manual V 2.0


Restricted

A7.1.1 Introduction
A typical API separator consists of two basic sections, the inlet section and the oil-water separation channels.
The body of the separator is generally constructed from concrete. Figure A7.1.1 illustrates a typical design.

Inlet section
The inlet section of the API separator consists of a pre-separator flume, trash rack, oil skimmer and forebay.

The pre-separator flume has two functions, the reduction of flow velocity and the collection of easily separated
hydrocarbons. Velocities in the flume are in the order of 0.05 to 0.10 m/sec, with a 1 to 2 minute residence
time.

The trash rack is a coarse screen used to remove sticks, rags, stones and other debris that would
interfere with the downstream equipment. It typically consists of inclined bars spaced on 25 mm to 50 mm
centres.

Easily separated hydrocarbons will collect on the surface of the water in the inlet section and will be retained
by a retention baffle. Skimming equipment is provided upstream of the retention baffle for the removal of the
collected hydrocarbons.

Separation section
After the inlet section the water flows into the separator channels. Shut-off gateways are provided at the inlet
to each channel to allow individual channels to be isolated if required. The water passes through a velocity
head diffusion device which is designed to reduce flow turbulence and to distribute the flow evenly across
the separator channel. Vertical slot baffles are generally used as the diffusion devices.

Once in the separation channel, the flow is stable and allows the separation of hydrocarbons and solids
from the water according to Stokes law. Hydrocarbons collect on the surface of the channel where they
are retained by a retention baffle and removed by an appropriate skimming device.

Solids will settle to the bottom of the channel. A mechanical scraper is usually provided for moving the sludge
along the bottom of the separator channel to a sludge collection point.

Parallel plate interceptor


The parallel plate interceptor (PPI) is basically an API separator with the separation section fitted with long
cross flow parallel plates to improve the separation efficiency. The concept of the PPI eventually lead to the
development of the more compact and efficient Corrugated Plate Interceptor (CPI) design which superseded
both the API and the PPI variation of the API.

A7.1.2. Performance variables


The separation characteristics of an API separator are governed by Stokes law. The theory of Stokes law is
discussed in Appendix A2.1. In general the API separator is designed for the removal of hydrocarbon drop-
lets larger than 150 mm.

A7.1.3. Installation/Configuration
The API separator is a primary separation device. It may be used as the first stage in a two or three
stage water treatment plant, or for the treatment of large water flows with easily separated hydrocarbons
such as drains water or ballast waters.

The size and weight of the API separator design prevents its use offshore.

Deoiling Manual V 2.0 129


Restricted

Figure A7.1.1: API separator with two separation channels and no skimming facilities on the inlet section.

A number of skimming devices are available for the removal of collected hydrocarbons from the surface of the
water including rotary drums, slotted pipe skimmers, horseshoe floating skimmers and self adjusting floating
skimmers. These are discussed in detail in the referenced guidelines.

A modification to the rotating drum type skimmer is the use of a oleophilic film on the surface of the drum. This
makes the drum more efficient at hydrocarbon skimming while minimising the quantity of water carried over
with the recovered hydrocarbons. The system can be fitted to existing API separators.

A7.1.4. Design guidelines


The original API manual gave the following guidelines for the sizing of the separation channel for an API
separator.
Maximum horizontal velocity in the separation channel is 0.015 m/sec or 15 times the hydrocarbon droplet
rising velocity (whichever is smaller).
The dimensions of the separator channel should fall within the following ranges.
Depth (metres): 0.9 < d < 2.4
Width (metres): 1.8 < w < 6.0
Ratio: 0.3 < d/w < 0.5

Unless provision is made for continuous sludge removal, an allowance of 0.3 meters should be made for the
accumulation of sediment on the bottom of the separator channel.

130 Deoiling Manual V 2.0


Restricted

Chapter 8.

Deoiling equipment –
dissolved hydrocarbons

Deoiling Manual V 2.0 131


Restricted

8. Dissolved hydrocarbons
8.1. Introduction
in many E&P operations, particularly offshore, both regulations governing hydrocarbon discharges and deoil-
ing equipment design are based on the removal of only insoluble dispersed hydrocarbons from waste water.

However, as illustrated in Chapter 2, waste water from hydrocarbon production facilities may contain sig-
nificant quantities of dissolved hydrocarbons. In many land based installations, particularly in the refining
and manufacturing industries, discharge regulations and environmental considerations frequently require the
removal of both dispersed and dissolved hydrocarbons from effluent water. In the future similar standards may
be applied to the discharge water from E&P operations, including offshore installations.

The currently preferred and practised technology for the removal of dissolved hydrocarbons in E&P operations
is Macro Porous Polymer Extraction (MPPE). This will be dealt with in detail in section 8.3.

8.2. Levels of dissolved hydrocarbons in water


The actual dissolved hydrocarbon concentration is a key parameter for the design of equipment for the re-
moval of dissolved hydrocarbons, yet is often not known at the design stage. A number of methods can be
used to determine the probable dissolved hydrocarbon concentration.
■■ Water samples taken during well tests can be analysed to determine the dissolved hydrocarbon content.
This can provide valuable information for subsequent design. The well test operating conditions should be
clearly noted.
■■ From hydrocarbon and water analyses, computer simulation may be used to predict expected the levels of
dissolved hydrocarbons. The accuracy of such prediction is typically in the order of ± 30%.
■■ Samples of produced hydrocarbons and produced water can be used in laboratory trials to investigate the
levels of dissolved hydrocarbons. While effects such as ageing will reduce the accuracy of laboratory tests,
they can be used to give broad confirmation of other prediction methods.

8.3. Macro Porous Polymer Extraction


8.3.1 General
The MPPE technology was developed in the early 1990’s by Akzo Nobel and was designed to remove
dissolved and dispersed hydrocarbons in water. The MPPE process has developed into a proprietary technol-
ogy, now supplied by Veolia Water Solutions & Technology (VWS). An annual fee needs to be paid to VWS
for the MPPE ongoing Performance Guarantee and related Services.
The MPPE process is based on liquid-liquid extraction, where the extraction liquid is immobilized in a Macro
Porous Polymer. The porous polymer particles have a diameter of 1 mm, with pore sizes of 0.1 – 10 µm.
The porosity is 70 to 80%. The polymer and extraction solvent are both highly hydrophobic, and therefore
remove the non-polar hydrocarbons from the water. This technology is capable of removing compounds with
a lower polarity than water down to ppb level, or as specified.

Periodical in situ regeneration of the extraction liquid is required by means of low pressure steam. The ex-
tracted hydrocarbons can be almost fully recovered.
Due to the regeneration process, the MPPE process is designed with two-columns to allow continuous opera-
tion, i.e. one column in regeneration mode and one column is in extraction mode.

132 Deoiling Manual V 2.0


Restricted

8.3.2. Process description


A schematic overview of the MPPE process is shown in Figure 8.1. The MPPE unit consists of two (2) columns,
both containing a packed bed of MPPE material. The influent water is fed into the first column of each unit

Figure 8.1: Schematic Overview of Macro Porous Polymer Extraction Process.

where the extraction process takes place in order to remove the contaminants. At the same time the second
column is being regenerated with low-pressure steam. After a pre-set time (usually approximately one hour),
the feed is switched to the second column for extraction. The first column is then regenerated by low-pressure
steam. After a period of time, the unit switches back again from the second to the first unit. The steam evapo-
rates the components from the MPPE material resulting in a vapour flow of organics and steam. The vapour
is routed through a condenser where condensation of both steam and organics takes place. The condensed
steam and organics are led into a separator in which the organics are separated from the condensed steam.
The practically 100 % pure organics can be spiked in the condensate/crude or disposed according to regu-
lations. The water phase from the separator is recycled to the MPPE system.

8.3.3. Separation performance


Dissolved, and to some extent dispersed hydrocarbons, will be removed by the MPPE process.
In current running units, dissolved aromatics (especially benzene), PAHs, aliphatics and chlorinated hydrocar-
bons are removed for more than 99.9 %. Usually, the size of the unit is tailored to the required separation per-
formance. The components that can be removed with the MPPE system must have an affinity for the extraction
liquid compared to water in order to allow extraction. Removal efficiency will depend on polarity of the hy-
drocarbon, the affinity for the MPPE extraction liquid (partition coefficient), and volatility of the hydrocarbon.

Deoiling Manual V 2.0 133


Restricted

Dissolved hydrocarbons
Dissolved hydrocarbons that can be removed by the MPPE process are:
■■ Aromatics (BTEX)
■■ PAHs (Polyaromatic Hydrocarbons),
■■ NPD’s (naphtalenes, phenanthrenes, dibenzothiophenes)
■■ TEOX (Total Extractable Organics)
■■ Halogenated hydrocarbons
■■ And in principle all other hydrophobic compounds (e.g. CS2)

BTEX can be removed with an efficiency of >95% with influent levels up to at least 2000 ppm. Benzene is
typically the design governing component to be removed.

Naphtalene removal of >99%, and PAH removal of >93%+ can be obtained. MPPE claims even removal
efficiency of PAH >99% for commercial running units. Usually, the size of a unit is tailored to the required
separation performance.

Phenolic substances
Phenolic substances have a high polarity and are therefore difficult to remove. For the total group of alkyl
phenols, an average removal efficiency of 30% has been measured (only a few times). Especially the lower
phenols (phenol, and o-/m-/-p- cresol) are difficult to remove, and will not be removed at all when they are
in ionic form.

Organic acids
Organic acids such as acetic acid, propionic acid will not be removed.

MTBE/ETBE
The removal of MTBE and ETBE (Methyl/Ethyl tert-butyl Ether) is difficult due to the high solubility in water.
MTBE/ETBE can be a governing component for MPPE design.

Lower alcohols
Methanol, glycol (i.e. MEG), tertiary butyl alcohol (TBA) etc. flow through without harming the process.

Dispersed hydrocarbons/Aliphatics
Aliphatic compounds include alkanes, alkenes, alkynes, and can be cyclic, like cyclohexane , or acyclic, like
hexane. C10+ hydrocarbons will be mainly dispersed hydrocarbons.

Veolia always refers to total aliphatic removal efficiency (i.e. C10-C40). However, the fraction C20+ aliphatic
components will be an important factor in the total aliphatic removal efficiency (C10-C40). Therefore, for
aliphatics/dispersed hydrocarbon removal it is important to make a distinction between gas/condensate
produced water, and oil produced water. The MPPE process is mainly applied in the water treatment proc-
ess of gas produced water. Removal efficiencies available in the public domain typically refer to aliphatics
removal (or dispersed hydrocarbon) removal in gas/condensate produced water.

More detailed performance data still need to be established.

Gas produced water


The aliphatic removal from a gas/condensate produced water at an undisclosed gas/condensate platform,
showed an aliphatics removal efficiency >99.5%. These aliphatics will be mostly <C20 aliphatics, as it
comes from a gas/ condensate platform. The exact fraction dissolved or dispersed aliphatics is unknown.

Furthermore, Veolia claims that 99% removal for aliphatics is achieved for Ormen Lange as well.

134 Deoiling Manual V 2.0


Restricted

Oil produced water


MPPE is less effective for dispersed hydrocarbon removal from produced water from oil fields.
■■ Aliphatics below C20 can be removed at 95% levels, with influent levels at hundreds of ppm.
■■ Aliphatics > C20 removal is less than 10% (for long term, steady-state operation).
The overall removal of dispersed hydrocarbons will depend on the size of the < C20 fraction .

The reduced removal of C20+ molecules is not a step change, as the effect is that the bigger the molecule,
the higher affinity to the MPPE resin, but also the lower the vapour pressure and therefore the less removal in
the regeneration step. The larger aliphatics will build up, and reach a steady state. Finally a breakthrough
will occur of the larger aliphatic molecules.

Veolia always emphasizes that the aliphatic C20+ will not impact aromatics, PAH removal.

At recent pilot testing of MPPE at Shearwater (2009), the dispersed HC removal failed completely. This is
almost certainly due to the fact that the inlet OIW concentration at Shearwater is relatively high (> 1000
ppm), and the droplets are exceptionally small and contaminated with a high concentration of surfactants.

8.3.4 Operating envelope


Capacity
The hydraulic capacity, which can be handled, based on the main equipment, is maximum 190 m3/h with
1 units in operation. Multiple units can be installed in parallel to obtain higher capacity.
The turndown/turn up ratio of a unit is 0 to approx 125 %.
Capacity can be tuned to the changing influent levels to obtain the required effluent, e.g. in case of a 50 %
higher influent concentration the same effluent level can be obtained with only a 10 % lower flow. At lower
influent concentrations, higher flow rates can be reached.

Suspended solids
Due to the high porosity of the bed (around 50 vol.%) most suspended particles will pass the bed and leave
the system with the effluent. However, to avoid accumulation of small particles, MPPE units are provided with
a pre-filtration step in the form of candle filters with 5-20 µm elements. Acceptable solid content in the feed
of cartridge filters is typical 5 mg/l or lower.

Oxygen
The MPPE process can operate under aerobic and anaerobic conditions.

Table 8.1 - Normal operation range for MPPE process.

Parameter Value Comment


Inlet pressure @ inlet 3-4 barg DP over MPPE unit ~ 3 barg:
Filtration: DP= 1 bar;
Control valve: DP= 1 bar;
MPPE bed: DP= 1 bar.
Standard design pressure for the MPPE col-
umns is 6 barg/FV
Max. feed temperature 55 °C to 60°C In case of benzene removal, the max recom-
mended operating temperature is 40 °C, as
10%-30% reduction can be expected at 50-
60°C for benzene.
Effluent temperature = feed temperature +3 to 4 °C Due to steam generation process
pH 3-9 pH can impact removal efficiency of alcohols

Deoiling Manual V 2.0 135


Restricted

Table 8.2 – Steam conditions for MPPE regeneration process.

Parameter Value Comment


Steam temperature Average ~110 °C Max temperature due to irreversible damage to MPPE
Max temperature <130 °C resin
Steam pressure 0.5 barg

8.3.5. Design guidelines


The standard design pressure for the columns is 6 barg/FV.
The largest MPPE unit ever built treats 120 m3/h of water. While scaling-up the MPPE unit, the upward liquid
velocity in the bed will be held constant. Hence, the diameter will increase with increasing flow rates. The
maximum diameter is 2.4 m for construction reasons. At higher throughputs, multiple parallel units will be
installed. The biggest diameter ever built is 2.2 m.
Velocity (empty bed) is in the range of 32 – 43 m/hr.

Duplex is preferred Material of Construction, as duplex is known for its resistance against elevated chloride
levels, and temperatures (to 130 °C). Duplex has been applied on several MPPE systems already. Other
material can be considered in consultation with the materials engineer.

The entire MPPE unit is classified as Zone 2.

8.3.6. Process Location


On a gas platform, the MPPE unit is typically located downstream the degasser.
In case of MEG regeneration unit, the MPPE can treat the produced water from the MEG regeneration unit to
reduce aromatic hydrocarbons (specifically benzene) to such a level that biotreatment is feasible.

8.3.7 Operational requirements


Utilities
The following utilities shall be provided from the existing utility system:
■■ Electricity: 380 – 415 V, max. load 50 kW;
■■ instrument air;
■■ LP steam @ 0.5-1.0 barg or water for steam generator. Typically, the system requires 3.5 to 5.9 kg water
per cubic meter treated water for steam production.
■■ Cooling water (which has low hardness).

Vent gas treatment


There are two locations within MPPE unit where organics are in contact with air:
■■ Separator vessels S-1X1;
■■ Return water tanks T-1X0.
Any level change in these tanks, especially an increase in level will lead to exhaust of vapour loaded air.
For the return water tanks this is regarded negligible due to the low concentration of organics in the water.
Especially when MPPE unit is designed for benzene removal, then vents need to be connected to a flare
system, or off gas treatment (i.e. AC filters) need to be installed

Exchange of MPPE material


Exchanging MPPE material is done by Veolia, as part of the service contract (PSG). Typically the resin needs
to be replaced once every two years.
Filling of new material is carried out by slowly emptying a big bag of MPPE material hanging over the col-
umn. The refilled column can directly be started again. The maximum expected downtime of one unit due to
replacement and refilling of the material is expected to be 3 days.
Sufficient access shall be allowed for the easy loading and unloading of the unit internals, e.g. material, resin
beds. MPPE material handling area is indicated on the plot plan.

136 Deoiling Manual V 2.0


Restricted

8.3.8. Veolia Services


The MPPE process is a proprietary process supplied by Veolia Water Systems. An annual fee needs to be paid
for the MPPE ongoing Performance Guarantee and related Services, i.e. MPPE material supply and exchange,
operations support on and off site (via remote control). Service fee ranges from 0.17-0.25 euro/m3, depending
on contaminants removal efficiency and capacity/throughput.

8.3.9. Operating experience


The following MPPE references are known within Shell:
■■ NAM K15A (since 2002, treating 3 – 5 m3/h);
■■ NAM K15B (since 2003, treating 3 – 5 m3/h);
■■ Shell-Ormen Lange gas condensate (start-up in 2007, 68 m3/h);

The first commercial installation was installed


at ELF-Petroland (now Vermillion) in Harlin-
gen, where a unit is in operation since 1994
(see Figure 8.2). Offshore gas produced
water is treated onshore via an MPPE unit at
a throughput of 3 – 6 m3/hr. The average
BTEX concentration is reduced from 1,500
(with peaks up to 3,000) down to 0.5 mg/l.

The largest unit currently in operation, has


been built in the Ruhr area (undisclosed end
customer, Germany) to treat 120 m3/h of
polluted ground water, containing chlorin-
ated hydrocarbons; 600 ppb inlet and <10
ppb at outlet). Start-up: 1998.

Figure 8.2 – MPPE Unit at Vermillion (Total) Harlingen. D=0.8m,

Deoiling Manual V 2.0 137

Potrebbero piacerti anche