You are on page 1of 51

Eni S.p.A.

Exploration & Production Division

DESIGN CRITERIA

INTERNAL CORROSION - FLUIDS CLASSIFICATION


AND CORROSION PARAMETERS DEFINITION

02555.VAR.COR.PRG
Rev. 3 – May 2006

ENGINEERING COMPANY STANDARD

Documento riservato di proprietà di Eni S.p.A. Divisione Agip. Esso non sarà mostrato a Terzi né utilizzato per scopi diversi da quelli per i quali è stato inviato.
This document is property of Eni S.p.A. Divisione Agip. It shall neither be shown to Third Parties not used for purposes other than those for which it has been sent.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 2 of 51

FOREWORD

Rev. 3 May 2006


ISSUE
N. Sheets 51

Rev. 2

No. Sheets 47
June 1995

The type of document has been changed from “GENERAL SPECIFICATION” to “DESIGN
CRITERIA”.

The Normative References chapter has been revised and updated.

It has been deeply revised the paragraphs concerning the corrosion parameters and the corrosion
forms; corrosion evaluation criteria have also been updated.

With respect to the previous revision, it has been completely eliminated the classification by
corrosion environments. In this revision, fluids are defined only in terms of “type of fluid” and
“corrosivity class”.

A detailed list of definition has been introduced.

Definition of sour conditions in accordance to EFC - European Federation of Corrosion, has been
introduced.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 3 of 51

TABLE OF CONTENTS
INTERNAL CORROSION - FLUIDS CLASSIFICATION AND CORROSION PARAMETERS
DEFINITION
1. GENERAL
1.1 Scope
1.2 Reference normative
1.2.1 Company specifications
1.2.2 International standards
1.3 Definitions
1.4 Symbols and abbreviations
1.5 Conversion factors
2. FLUID TYPES
2.1 Liquid hydrocarbons and multiphase
2.2 Gas and gas with condensate
2.3 Waters
2.4 Glycol
2.5 Amine
3. CORROSION PARAMETERS
3.1 Temperature
3.2 Pressure
3.3 Water cut
3.4 Water wetting
3.4.1 Liquid and multiphase systems
3.4.2 Gas and gas with condensate systems
3.5 Gas oil ratio / Condensate gas ratio / Gas liquid ratio
3.6 Hydrodynamic conditions
3.6.1 Flow pattern
3.7 CO2 and H2S molar fraction and partial pressure
3.8 API grade
3.9 Water Chemistry
3.10 Bacteria
3.11 Sand and solid particles
3.12 Elemental sulfur
3.13 Mercury
4. CORROSIVITY CLASSES
4.1 Hydrocarbon systems and production water
4.1.1 Corrosion parameters
4.1.2 Corrosivity classes
4.2 Service waters
4.2.1 Corrosion parameters
4.2.2 Corrosivity classes
4.3 Glycol
4.3.1 Corrosion parameters
4.3.2 Corrosivity classes
4.4 Amine
4.4.1 Corrosion parameters
4.4.2 Corrosivity classes
5. CORROSION FORMS
5.1 General
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 4 of 51

5.1.1 Foreword
5.1.2 Materials
5.1.3 Corrosion morphologies
5.2 CO2 corrosion
5.2.1 Effects of chemical species in solution
5.2.2 Effect of H2S on CO2 corrosion
5.2.3 CO2 corrosion prediction models
5.2.3.1 NORSOK M-506 CO2 corrosion rate calculation model
5.2.3.2 De Waard & Milliams empirical model (1991/1993)
5.2.3.3 De Waard, Lotz, & Dugstadt semi-empirical model (1995)
5.2.3.4 Effects of corrosion inhibition treatment and glycol
5.2.4 Top of line corrosion
5.2.5 Corrosion products
5.3 Sour Service
5.3.1 Sulfide Stress Cracking (SSC)
5.3.2 Sour Service Resistant Materials
5.3.3 Hydrogen Induced Cracking (HIC)
5.3.4 H2S corrosion
5.4 Pitting and crevice
5.4.1 Initiation and propagation conditions
5.5 Stress Corrosion Cracking (SCC)
5.5.1 Chloride Stress Corrosion Cracking
5.5.2 Polythionic Acid Stress Corrosion Cracking
5.5.3 Amine (or Alkaline) Stress Corrosion Cracking
5.6 Oxygen corrosion
5.6.1 Carbon and low alloy steels. Aerated water
5.6.2 Carbon and low alloy steels. Deaerated water
5.6.3 Corrosion resistant alloys
5.7 Erosion-corrosion and erosive wear
5.7.1 Totally solid-free systems
5.7.2 Nominally solid-free systems
5.7.3 Solid particles laden systems
5.8 Microbial Induced Corrosion (MIC)
5.9 Galvanic corrosion
5.10 Elemental sulfur corrosion
5.10.1 Elemental sulfur from the reservoir
5.11 Amine corrosion
5.11.1 Carbon steels
5.11.2 Austenitic stainless steels
6. BIBLIOGRAPHY
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 5 of 51

1. GENERAL

1.1 Scope

This document provides criteria to classify, from a corrosion perspective, the main fluids encountered
in the upstream oil and gas production plants.

Fluid systems herein covered are relevant to process and utility facilities from the wellhead battery
limit (well is excluded) to the delivery of treated gas and stabilised oil. Refining plants are outside the
scope of this document.

For each type of fluid the document lists the parameters to be collected to allow the corrosion expert
assessment of the fluid corrosivity and to select fit-for-purpose materials and corrosion control
methods.

The most common corrosion forms in oil and gas industry are reviewed, providing guidelines for
corrosion prediction.

The following aspects are outside the scope of this document:


− external corrosion (atmospheric, soil or sea water)
− non metallic materials (elastomers, ceramics, composites, coatings).

1.2 Reference normative

1.2.1 Company specifications

03587.MAT.COR.PRG Metallic materials in contact with H2S containing environments.


Corrosion tests methods and evaluation criteria.

05489.MAT.COR.SDS Additional requirements for pressure vessels for applications in H2S


containing environments.

1.2.2 International standards

ISO 8044 Basic Terms and Definitions on Corrosion

ISO 151561 / NACE MR0175 Petroleum and natural gas industries - Materials for use in H2S-
containing environments in oil and gas production.
Part 1 – General principles for selection of cracking-resistant
materials.
Part 2 – Cracking-resistant carbon and low alloy steels, and use of
cast irons.
Part 3 – Cracking-resistant CRAs (corrosion resistant alloys) and
other alloys

1
ISO 15156 replaces NACE MR0175 “Metals for sulfide stress cracking and stress corrosion cracking resistance in sour
oilfield equipments” 2003 version.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 6 of 51

API RP-14E Design and Installation of Offshore Production Platform Piping


System

API 581 Risk-based Inspection Resource Document. Appendix G and H

API RP 945 Avoiding Environmental Cracking in Amine Units

API SPEC 12GDU Specification for Glycol-Type Gas Dehydration Units

ASTM 287-92 Standard Test Method for API Gravity of Crude Petroleum and
Petroleum Products (Hydrometer Method)

ASTM G 78-46 Standard Test Method for Pitting and Crevice Corrosion Resistance
of Stainless Steels and Related Alloys by the Use of Ferric Chloride
Solutions

NACE TM0177 Laboratory Testing of Metals for Resistance to Specific Forms of


Environmental Cracking in H2S Environments

NACE TM0284 Evaluation of Pipeline And Pressure Vessel Steels for Resistance to
Hydrogen Induced Cracking

NACE RP0472 Methods and Controls to Prevent In-Service Environmental Cracking


of Carbon Steel Weldments in Corrosive Petroleum Refining
Environments

NACE RP0296 Guidelines for Detection, Repair, and Mitigation of Cracking of


Existing Petroleum Refinery Pressure Vessels in Wet H2S Service

ASME B31.3 Process Piping

NORSOK M-001 Materials Selection

NORSOK M-506 CO2 corrosion rate calculation model

EFC Publication No 16 Guidelines on materials requirements for carbon steel and low alloy
steels for H2S – Containing environments in Oil and Gas Production

EFC Publication No 17 Corrosion Resistant Alloys for Oil and Gas Production: Guidance on
General Requirements and Test Methods for H2S Service. The
Institute of Materials

EFC Publication No 23 Design Consideration for CO2 corrosion in Oil and Gas Production

DNV RP O501 Erosive Wear in Piping Systems

1.3 Definitions

Chloride Stress Corrosion Cracking - CSCC


Formation of cracks caused by stress corrosion in a water and chloride ions-containing
environments.

Corrosion
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 7 of 51

Physicochemical interaction between a metal and its environment that results in changes in the
properties of the metal and which may often lead to impairment of the function of the metal, the
environment, or the technical system, of which these form a part (ISO 8044).

Corrosion product
Substance formed as a result of corrosion (ISO 8044).

Corrosion rate
Corrosion effect on a metal per unit of time (ISO 8044).

Corrosion resistance
Ability of a metal to withstand corrosion in a given corrosion system (ISO 8044).

Corrosion system
System consisting of one or more metals and all parts of the environment which influence corrosion
(ISO 8044).

Corrosive agent
Substance which when in contact with a given metal will react with it (ISO 8044).

Corrosive environment
Environment that contains one or more corrosive agent (ISO 8044).

Corrosivity
Ability of an environment to cause corrosion in a given corrosion system (ISO 8044).

Corrosivity class
In this document, it is an attribute conventionally assigned to each type of fluid in order to point out
the most significant corrosivity features.
For hydrocarbons (LH and GH) and production water (PW) the corrosivity class is assigned on the
basis of CO2 and H2S partial pressures as follows:
– N non containing CO2 and H2S
– C containing CO2
– CS containing CO2 and H2S
For the fluids designed as glycol (GL) and amine (AM) the corrosivity class is assigned as follows:
– L lean (free from water CO2 and H2S)
– R rich (containing water and possibly CO2/H2S)
For service waters (W) the corrosivity class is assigned as follows:
– D deaerated (oxygen taken off)
– A aerated (containing oxygen)

Crevice Corrosion
Corrosion associated with, and taking place in, or immediately around, a narrow aperture or
clearance (ISO 8044).

Dew point temperature


It is the temperature, below which liquid condensation starts from gas phase at a given pressure. On
the state diagram condensation conditions are indicated by the dew point curve. In particular the
water dew point refers to condensation conditions of water from gas.

Fugacity
It is a thermodynamic function, in pressure units, that, when used in a thermodynamic equation of an
ideal gas in substitution of pressure, allows to apply the same function to a non-ideal gas.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 8 of 51

Hydrocarbon
− Gas: a mixture of hydrocarbons with 1 to 4 carbon atoms at a temperature above the critical
temperature. Gas can be in form of dry gas or gas with condensates depending on
thermodynamic conditions.
− Liquid: a mixture of hydrocarbons whose temperature is below the critical temperature of the
particular system of natural hydrocarbons that the mixture contains; in the mixture the liquid
phase, always present, can be combined with a gas phase or an aqueous phase or both; in this
cases the system is called multiphase.

Hydrogen embrittlement
A process resulting in a decrease of the toughness or ductility of a metal due to absorption of
hydrogen. (ISO 8044).

Hydrogen Induced Cracking - HIC


Planar cracking that occurs in carbon and low alloy steels when atomic hydrogen diffuses into the
steel and then combines to form molecular hydrogen at trap sites. Cracking results from the
pressurization of trap sites by hydrogen. No externally applied stress is needed for the formation of
hydrogen-induced cracks. Trap sites capable of causing HIC are commonly found in steels with high
impurity levels that have a high density of planar inclusions and/or regions of anomalous
microstructure (e.g. banding) produced by segregation of impurity and alloying elements in the steel.
This form of hydrogen-induced cracking is not related to welding (ISO 15156).

Hydrogen Stress Cracking - HSC


Cracking that results form the presence of hydrogen in a metal and tensile stress, residual and/or
applied (ISO 15156).

Microbial Corrosion
Corrosion associated with the action of micro-organisms present in the corrosion system (ISO 8044).

Molar fraction
Measurement of the concentration of a chemical species expressed as ratio between the number of
moles of the given chemical species and the total number of moles.

Oil
In this document the terms oil and crude are considered synonymous to indicate liquid hydrocarbons.

Passivation
Decrease of corrosion rate by formation of a passive layer, i.e. a thin, adherent, protective layer
formed on a metal surface through reaction between metal and environment (ISO 8044).

Passivity (passive state)


State of a metal resulting from its passivation (ISO 8044)..

Pitting Corrosion
Corrosion resulting in pits, i.e. cavities extending from the surface into the metal (ISO 8044).

Predicted corrosion rate


It is the corrosion rate, usually expressed quantitatively (in mm/y) and/or qualitatively, determined: (a)
after the corrosion study, applying all the available knowledge and tools; (b) through laboratory tests,
simulating the real conditions; (c) on the base of field corrosion monitoring data applicable to the
case under study. The following categories are recommended to express in a qualitative way the
penetration rate for general corrosion forms: negligible, low, moderate, severe and very severe.

Reservoir
− Dry gas: reservoir whose temperature is above the cricondetermical temperature of the particular
system of natural hydrocarbons that it contains. The cricondetermical temperature is the highest
temperature at which the coexistence between liquid and gas phase is still possible.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 9 of 51

− Gas with condensates: reservoir whose temperature is between critical and cricondetermical
temperatures of the particular system of natural hydrocarbons that it contains. The condensation
of the liquid phase from gas takes place by reverse condensation, that is the phenomenon by
which decreasing the pressure below the dew point, there is initially an increase of the liquid
phase percentage and eventually a partial or total ri-evaporation of the latter.
− Crude oil: reservoir whose temperature is below the critical temperature of the particular system
of hydrocarbons that in contains.

Residual corrosion rate


It is the corrosion rate after treatments with corrosion inhibitors.

Sour Service
Exposure to oilfield environments that contain H2S and can cause cracking of materials by the
mechanisms addressed by ISO 15156 (ISO 15156).

Specific gravity
For a liquid it is the ratio between the weight of a given volume of liquid and the weight of the same
volume of water. For a gas it is the ratio between the weight of a given volume of gas and the weight
of the same volume of dry air in the same standard conditions.

Stepwise Cracking - SWC


Cracking that connects hydrogen-induced cracks on adjacent planes in a steel. This term describes
the crack appearance. The linking of hydrogen-induced cracks to produce stepwise cracking is
dependent upon local strain between the cracks and embrittlement of the surrounding steel by
dissolved hydrogen. HIC/SWC is usually associated with low-strength plate steels used in the
production of pipes and vessels (ISO 15156).

Stress Corrosion Cracking - SCC


Cracking of metal involving anodic processes of localized corrosion and tensile stress (residual
and/or applied) in the presence of water and H2S. Chlorides and/or oxidants and elevated
temperature can increase the susceptibility of metals to this mechanism of attack (ISO 15156).

Sulfide Stress Cracking - SSC


Cracking of metal involving corrosion and tensile stress (residual and/or applied) in the presence of
water and H2S. SSC is a form of hydrogen stress cracking (HSC) and involves embrittlement of the
metal by atomic hydrogen that is produced by acid corrosion on the metal surface. Hydrogen uptake
is promoted in the presence of sulfides. The atomic hydrogen can diffuse into the metal, reduce
ductility and increase susceptibility to cracking. High strength metallic materials and hard weld zones
are prone to SSC (ISO 15156).

1.4 Symbols and abbreviations

API Oil API Grade


CLR Crack Length Ratio
C Erosion Corrosion Constant
CA Corrosion Allowance
CE Carbon Equivalent
CGR Condensate Gas Ratio
CIE Corrosion Inhibitor Efficiency
CR Predicted Corrosion Rate
CRA Corrosion Resistant Alloy
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 10 of 51

CS Carbon Steel
CSCC Chloride Stress Corrosion Cracking
CSR Crack Surface Ratio
CTR Crack Thickness Ratio
DL Design Life
DWM deWaard & Milliams Model
FCOND deWaard Model Condensation Factor
FpH deWaard Model pH Factor
FSCALE deWaard Model Scaling Factor
FSYSTEM deWaard Model System Pressure Factor
FWW Water Wetting Factor
GLR Gas Liquid Ratio
GOR Gas Oil Ratio
HIC Hydrogen Induced Cracking
HSC Hydrogen Stress Cracking
MIC Microbiological Induced Corrosion
P Pressure
pCO2 CO2 Partial Pressure in Gas Phase
PRE Pitting Resistance Equivalent Number
PWHT Post Weld Heat Treatment
SCC Stress Corrosion Cracking
SMYS Specified Minimum Yield Strength
SSC Sulfide Stress Cracking
SWC Step Wise Cracking
T Temperature
TDS Total Dissolved Solids
TSCALE deWaard & Milliams Scaling Temperature
UAVG (U) Average Flow Velocity
UC Erosion Critical Flow Velocity
USG Gas Superficial Velocity
USL Liquid Superficial Velocity
vCO2 Expected CO2 Corrosion Rate, calculated from vdWM,C corrosion rate and reduced
to take into account water wetting, FWW, and inhibitor efficiency, CIE.
wCUT Water Cut (water volume/total liquid volume)
xCO2 CO2 Molar Fraction
xH2S H2S Molar Fraction
LME Liquid Metal Embrittlement
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 11 of 51

1.5 Conversion factors

bar = 14.5 psi


pound (lb) = 0.454 kg
mpy = 0.0254 mm/y
ksi = 6.895 MPa
°F = 1.8 × °C + 32
bbl (oil) = 0.159 m3
lb/ft3 = 16.02 kg/m3
ft = 0.3048 m
m3/m3 = 5.62 ft3/bbl
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 12 of 51

2. FLUID TYPES

This document covers the following types of fluids:

Fluid types Abbreviation


liquid hydrocarbon and multiphase systems LH
gas hydrocarbon and gas with condensates GH
production water PW
glycol GL
amine AM
service waters
− sea water SW
− fresh water FW
− brackish water BW

The above listed fluids represent the main categories met in the upstream oil and gas production
plants with a corrosive potential for metallic materials. Fluids met in the refinery industry such as
fluids for decarbonation and desulfuration treatment are excluded, as well as all chemical additives
(corrosion inhibitors, fluidizers, etc.).

2.1 Liquid hydrocarbons and multiphase

These fluids include oil, alone or in combination with gas hydrocarbons and/or water.

2.2 Gas and gas with condensate

These fluids include gas, alone or in combination with condensates or water. Reference is
particularly made to fluids coming from gas reservoirs.

2.3 Waters

In the upstream oil and gas production and utility facilities, several types of water may be involved,
with different characteristics corrosion wise. The main distinction is between production water and
service waters.

Production water is oxygen-free water formation or condensed water coming from the reservoir in
combination with produced hydrocarbon. For corrosion purposes, production water is classified and
assessed with the same approach used for hydrocarbons. If production water is used for water
injection into the formation, adequate corrosion control treatments are normally performed consisting
of biocide injection and physical or chemical deaeration to remove any contamination with oxygen.

Service waters are used for several purposes for oil and gas production and utilities. Waters are
classified on the basis of their origin as follows:
– Fresh water is usually desalinated shallow water (from rivers or lakes) or water from shallow
formations with low salinity. In general salinity is below 2 g/l.
– Brackish water is shallow water with a higher salinity approximately around 5 g/l and in general
above 2 g/l.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 13 of 51

– Sea water: has an average salinity of 35 g/l, but locally can assume very different values. When
used for water injection into the formation an adequate de-aeration treatment is usually
performed.

The following criteria are used to classify waters:

Production water - PW
− reservoir (formation and/or condensed) water

Fresh water - FW
− shallow waters
− TDS < 2.0 g/l

Brackish water - BW
− shallow waters
− TDS > 2.0 g/l

Sea water - SW
− TDS ≅ 35 g/l

2.4 Glycol

Reference is made to pure glycol or water/glycol mixtures. In the upstream oil and gas process,
glycol is commonly used:
− to prevent hydrates formation by injecting it into the hydrocarbon streams;
− for gas dehydration purposes.

In the first case, glycol if beneficial for corrosion control, by increasing pH and keeping water in
solution. Methanol is also used as an alternative to glycol for hydrate prevention with similar effects
on corrosion, but with more scattered results in terms of corrosion reduction.

In the second case, dry glycol (lean) flows against the gas stream into the gas dehydration column.
Glycol outcoming from the dehydration column is denominated ‘rich’ and contains water and, if
present, acid gases. Water is normally in solution with glycol, which greatly reduces actual
corrosivity. Corrosion prediction models as Norsok M-506 or deWaard & Milliams may be used to
evaluate CO2 corrosion rate. Indicatively, corrosion rate of carbon steel is generally very low when
pCO2 < 0.2 bar, requires pH control (buffering) or corrosion inhibitor injection when 0.2 < pCO2 < 2
bar and requires design or operational corrosion control when pCO2 > 2 bar, but also oxygen
contamination (dissolved in the glycol), H2S, chlorides intrusion and pH value are important
influencing factors. Also glycol degradation products such as organic acids may lower the pH leading
to a corrosive environment. Corrosion control methods usually consists in the control of the stream
quality, pH control with buffering or neutralisers and filtration to remove solids (to avoid under deposit
corrosion). pH in the range 1 ÷ 8 are usually satisfactory (API SPEC 12GDU).

Materials to be used for H2S-containing glycol shall be compliant with sour service requirements as
addressed into ISO 15156 / NACE MR0175 for sour service.

Rich (wet) glycol is regenerated to remove water and acid gases and re-circulated back to
dehydration unit as lean glycol. Lean glycol is not corrosive.

2.5 Amine

Reference is made to mixtures of water and amines. Amines in the upstream oil and gas process are
commonly used for sweetening purposes.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 14 of 51

Monoethanolamine (MEA), diethanolamine (DEA) and methildiethanolamine (MDEA) solvents are


used to remove acid gases, primarily H2S and CO2, from plant streams. Diisopropanolamine (DIPA)
and diglycolamine (DGA) are also used in some treating plants.

MDEA offers process advantages and has become a major alternative to MEA and DEA. The most
important advantage is gas selectivity: MEA and DEA also remove CO2, but MDEA is selective to
H2S and removes little CO2 if present. Other MDEA advantages are energy saving and the ability to
operate at higher concentrations than MEA and DEA.

Rich amine contains acid gas (CO2/H2S) and involves corrosion concerns in terms of amine stress
corrosion cracking for low alloy and carbon steels (see parapgraph 5.5.3), general corrosion and
localised corrosion for carbon steels and austenitic stainless steels (see paragraph 5.11).

Materials to be used with H2S-containing fluids shall be compliant with sour service requirements as
addressed into ISO 15156 / NACE MR0175 for sour service.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 15 of 51

3. CORROSION PARAMETERS

Corrosivity assessment of the fluids is carried out on the basis of a set of parameters which include
both fluid data and operating conditions. The following paragraphs discuss the main affecting
parameters to be considered.

3.1 Temperature

Temperature has complex effects on corrosion rate, affecting the aggressiveness of a fluid in
different ways. Operating temperature values ‘T’ shall be collected, as well as maximum fluid
temperature ‘TMAX’ and minimum temperature ‘TMIN’ that the system or the examined component can
experience, including ambient temperature, transitory and upset conditions.

3.2 Pressure

Pressure is of primary importance for corrosion. Primarily, it determines the value of the partial
pressure of corroding agents as CO2, H2S, O2 into a gas system which define, through the Henry’s
law, the respective contents in the water phase in equilibrium with the gas.

3.3 Water cut

Water is the carrier of corrosion therefore water content in production fluids is a key factor for
corrosion assessment. Further than its presence as a separate and liquid phase, the condition for
corrosion to occur involves the ‘water wetting’ of the metallic surface exposed to the fluid.

In liquid hydrocarbon or multiphase system, water wetting depends on the water content, the oil
capability to entrap water into emulsion, the flow velocity and the flow pattern.

Water content in liquid and multiphase systems is expressed as water cut ‘WCUT’, in m³/m³ %, that is
the percent ratio of water volume on total volume of liquid phases:

%H 2O
W CUT =
% H 2 O + % LH

In production hydrocarbons systems, the associated water can be originated either by condensation
from wet gas or it can be dragged from the reservoir as formation water. In the latter case water is
usually characterized by a high salinity and a higher pH.

Gas from the reservoir is usually associated with water condensing from the gas itself, unless the gas
temperature is at least 10 °C above its dew point temperature, in which case the gas is dry and not
corrosive. In wet gas systems, water wetting from condensing water is always anticipated.

3.4 Water wetting

Corrosion can occurs when a separated and free liquid water phase is present and is in permanent
or temporary direct contact with the metallic surface. These conditions define the water wetting
occurrence. The corrosion attack is proportional to the fraction of time water wetting conditions are
verified.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 16 of 51

Ionic species in water solutions take action as reagents, products or intermediates, in the
electrochemical corrosion reactions. Water activates local micro cells on metallic surfaces. In specific
conditions, other solvents such as methanol, can perform a function similar to water.

3.4.1 Liquid and multiphase systems

In multiphase systems corrosion is proportional to the fraction of time the metal is microscopically
wetted by the aqueous phase. In multiphase systems, water separation and transportation of the
aqueous phase in contact with the metallic wall (wetting), depends primarily on the water cut and
secondarily on a complex interaction of factors, including:
− nature of the phases and repartition;
− phases composition/properties;
− hydrodynamic;
− geometry.

The water cut in hydrocarbon production is quite variable. When water cut is very low, below a few
percents, water is most likely completely emulsified into oil and the metallic walls are oil-wetted. On
the contrary, when water is the predominant phase, water wetting is expected. Intermediate
conditions are uncertain, and there is the possibility for intermittent or permanent water wetting at
specific locations.

Local water wetting conditions can occur in presence of geometrical discontinuities, which cause
turbulence and water separation from the oil phase, even if water cut is very low.

In stagnant or quasi-stagnant conditions (at the bottom of process vessel, tanks or in horizontal pipes
in laminar flow, with flow velocity lower than 1 m/s), water phase spontaneously separates and
stratifies at the bottom and water wetting occurs also when the water cut is extremely low, therefore
water wetting is always expected if water is present.

3.4.2 Gas and gas with condensate systems

In gas and gas with condensates reservoirs, gas is saturated with water, in equilibrium with the liquid
water in the reservoir. During production, the gas expands and cools: when it reaches the water dew
point, water starts to condensate from gas. Liquid gasoline condenses as well, following or
anticipating water condensation.

Dealing with wet gas or gas condensates, and without reliable information on the presence of water
and condensation dynamics, it is conservative to assume that an aqueous phase is present and
wetting the metallic surface. On the contrary, in gas systems operating at temperature at least 10°C
above their water dew point water condensation cannot occur. These systems are classified as “dry
gas” and are not corrosive as far as the temperature is maintained higher than the dew point.

Further than condensation water, also formation water dragged from the reservoir can be associated
to the production gas. In the case of condensed water, salinity is very low, whereas in formation
waters is typically high. An intermediate condition occurs when production water is made up of a
mixture of both. For corrosion assessment purposes, water with a total dissolved salinity higher than
1 g/l may be assumed to be of formation origin.

Depending on hydrodynamic conditions, water can exist in form of small drops dispersed in the gas
or as a liquid layer wetting the pipe walls while flowing. The thickness of the liquid layer is tens to
hundreds microns and can be composed by water, hydrocarbons or a mixture of both. In any case,
unless water presence is excluded, water wetting conditions are anticipated because liquid
condensate is poorly capable to emulsify water even at very low water cut.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 17 of 51

3.5 Gas oil ratio / Condensate gas ratio / Gas liquid ratio

Gas Oil Ratio ‘GOR’ and Condensate Gas Ratio “CGR” are the production ratios of relevant
hydrocarbon phases referred to standard (or normal) conditions, i.e. at 1 bar and 25 °C, and they are
expressed as std.m³(gas)/m³(oil) or std.m3(condensate)/std.m3(gas).

GOR usually varies during reservoir exploitation, typically increasing with time. High GOR are usually
characterised by a higher tendency to turbulent flow, increased water wetting and erosion-corrosion
likelihood, especially at flow disturbed flowing sections.

Gas Liquid Ratio ‘GLR’ is the produced gas to produced liquids (oil and water) ratio. It is referred to
standard (or normal) conditions, i.e. at 1 bar and 25 °C, and it is expressed as st.m³(gas)/m³(oil and
water).

3.6 Hydrodynamic conditions

For monophase flows, gas or liquids the average flow velocity or flow rate is defined as the ratio
between flow and cross section of the pipe.

In multiphase systems, it is defined for each phase a ‘surface flow velocity’ as the velocity calculated
considering each phase flowing alone in the full pipe cross section. Surface velocities in multiphase
systems are then used to estimate flow pattern by entering specific maps for horizontal or vertical
flows.

Hydrodynamic conditions affect corrosion in the following ways:


− effects on phase transportation: high turbulence conditions, particularly in correspondence of
geometrical discontinuities (for instance in correspondence of protruding welds), could cause
local water separation; on the other hand, low flow (typically below 1 m/s) or stagnant conditions
promote water separation from oil and stratification at the bottom of the pipe or vessel;
− effects on transport phenomena in solution: turbulence affects mass transport of corrosion
reactions reagents and products;
− decrease in corrosion inhibitor efficiency at high flow rate conditions (typically above 5 m/s for
liquid and multiphase systems);
− mechanical effects in removing protective deposits (corrosion products or scales) on metallic
surfaces;
− mechanical abrasion effects on metallic walls in sand or solid particles containing fluids.

These effects are quite complex to be investigated and worth involving corrosion and hydrodynamic
experts.

3.6.1 Flow pattern

The flow pattern depicts the repartition of the different immiscible phases while flowing through the
pipe section. The particular pattern established in a pipe depends mainly on flow rates, fluid
properties and tube sizes.

In monophase liquid oil/water systems the flow patterns in horizontal pipes are predicted on the basis
of the average velocity of the fluid in the pipe (UAVG) and the water cut (WCUT). Liquid system flow
patterns are indicatively identified as follows (Ref. /10/ and Ref. /12/):
– oil-in-water dispersed UAVG > 1 m/s & WCUT > 30%;
– water-in-oil dispersed UAVG > 1 m/s & WCUT < 30%;
– stratified UAVG < 1 m/s.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 18 of 51

In multiphase gas/liquid systems flowing in horizontal pipes, the superficial velocities are used. The
superficial velocity is defined as the velocity the individual phase (USL and USG respectively) would
exhibit flowing alone through the whole pipe cross section (USX = QX/A).

Simplified rules for flow pattern prediction in horizontal multiphase systems are the following:
– stratified/wavy USL < 0.15 m/s & USG < 10 m/s;
– slug 0.15 < USL < 5 m/s and 1 < USG < 22 m/s;
– annular USL < 4 m/s and USG > 22 m/s.
– dispersed USL > 5 m/s or 0.15 < USL < 5 m/s and USG < 1 m/s

Maps for flow pattern prediction as a function of superficial liquid and superficial gas velocities are
available in the technical literature (Ref. /1/, Ref. /2/, Ref. /3/). The following figures represent
examples of maps for flow pattern prediction in horizontal multiphase flow.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 19 of 51

3.7 CO2 and H2S molar fraction and partial pressure

When CO2 or H2S or both are present in the gas phase, at a content expressed as molar fraction
‘xCO2/xH2S’, any water in contact with this gas dissolve CO2/H2S to a concentration proportional to
the CO2 partial pressure in the gas ‘pCO2’ in accordance to the Henry’s law:
[CO2]L = HCO2 pCO2 ; [H2S]L = HH2S pH2S
The Henry constant H decrease with temperature and salinity.

The partial pressures in the gas phase, pCO2 and pH2S, are the main parameters respectively used
for CO2 corrosion rate evaluation of carbon steel and for sour service evaluation. Partial pressure are
given by the following expressions
pCO2 = P ⋅ xCO2 ; pH2S = P ⋅ xH2S
At high pressure, roughly above 100 bar (10 MPa), the effect of gas deviation from ideality becomes
significant and fugacity “f” should be used to calculate partial pressure:
fCO2 = a × pCO2 ; fH2S = a ⋅ xH2S
where fugacity coefficient “a” is given by the following expressions:
1.4
P×( 0.0031− )
a = 10 T + 273

For a wet gas or a multiphase system xCO2 or xH2S is the local CO2 or H2S molar fraction in the gas
phase, whereas for a liquid system it is the molar fraction in the last gas phase which was in
equilibrium with the liquid (e.g. the gas separator in the relevant outcoming liquid stream). In liquid
system, i.e. above the bubble point pressure, any increase in pressure gives no contribution to the
CO2 or H2S content in the water phase.

3.8 API grade

API grade, or API gravity, is a measure of fluid density for liquid hydrocarbons (ASTM D 287-92). It is
calculated with the following expression:
141 .5
°API = − 131 .5
γ
γ = specific gravity at 15 °C.

Liquid hydrocarbon display values ranging from 6° (heavy) to 60° (light). Usually oils range from 25°
to 35° API grade. Liquid hydrocarbons are considered light when they have API grade values in
between 35° and 45°, and heavy when API grade is below 25°.

Heavy oils often present a higher capability to promote water-in-oil emulsion, avoiding corrosion. On
the contrary, water separates easily from light gasoline and water wetting occurs even at very low
water cut.

3.9 Water Chemistry

To evaluate corrosivity properly, the chemical analysis of the water phase is required. Information
regarding sampling procedures shall be collected for a clear interpretation.

Some chemical composition parameters have primary importance and are reviewed in this section.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 20 of 51

Total salinity (TDS)

It is the dry residue at 110°C after water evaporation (boiling), expressed in g/l.

High salinity, especially in presence of oxygen, enhances localised corrosion and stress corrosion
cracking on susceptible materials, promoting separation between anodic and cathodic areas, and
galvanic coupling effects.

pH

The water phase in situ pH has a significant effect on corrosivity and it is of paramount importance
for all corrosion forms. In fact, pH influences the hydrogen evolution reaction, the precipitation
equilibrium, the nature and protectiveness of corrosion products, stress corrosion cracking and
localised corrosion.

In aqueous phase associated with hydrocarbon production, pH is associated to the solubilization


equilibria of acid gas as CO2 and H2S depending on the partial pressure and water phase
composition.

The pH value reported in water analyses is usually measured after depressurisation and atmospheric
exposure and give no information about the real in situ pH to be used for corrosion rate evaluation.
The pH direct measurement at operating conditions, in situ pH, is not practical and its value and can
be determined by calculations which take into account the three controlling buffer systems:
CO2/HCO3-, H2S/HS- and CH3COOH/CH3COO-. The buffer system CH3COOH/CH3COO- can be
extended to include all the other organic acids that may influence the pH determination as these
acids have similar dissociation constants.

Software programs, based on temperature, pressure and water composition are available for its
estimation. A simple approach is proposed by ISO 15156 - Annex D (see next figures): and is based
on temperature, bicarbonates, CO2 and H2S partial pressure in the associated gas phase. An
alternative approach based on equilibrium constants is provide by Norsok M-506.

CONDENSED WATER FORMATION WATER


Extract from ISO 15156-2 Annex D – pH determination

The bicarbonate concentration, which is used for pH calculation, is routinely determined by titration
and the reported value is set equal to the alkalinity. This is correct if H2S concentration is low and the
system does not contain acetic acid. For waters containing significant amount of organic acids, the
reported bicarbonate concentration requires correction because organic acids contributes to the
measured value of total alkalinity.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 21 of 51

Acid environments (pH < 6) are more corrosive than neutral (pH from 6 to 8).

Chlorides

Chlorides affect localised corrosion and stress corrosion cracking of materials with a passive
electrochemical behaviour, mainly through the depassivation effect caused by chloride ions,
especially with stainless steels. Chlorides, under specific circumstances, can also accelerate
corrosion of carbon steel by interacting with corrosion products and promoting localised corrosion,
especially in H2S dominated services where the corrosion product is FeS.

Oxygen

Water in equilibrium with the atmosphere dissolves oxygen. The solubility of oxygen in aerated water
decreases with temperature and salinity.

Reservoir fluids are oxygen-free. If the production fluid is contaminated with oxygen (atmospheric
exposure, faulty seals, as contaminant of additives etc.), the effect on fluid corrosivity can be severe
and a careful assessment is required. In particular for H2S-containing fluids, oxygen may cause H2S
oxidation to elemental sulfur, tiosulfates etc., which are very corrosive for most materials.

Bicarbonates

Bicarbonates (HCO3-) are usually present in formation water associated with production fluid. The
bicarbonates content influences water phase pH as above described (see pH section).

Organic acids

Organic acids (formic –HCOOH- and acetic -CH3COOH-) are often present in production fluids
containing CO2. Organic acids can be at the origin of enhanced corrosivity. Their presence shall also
be considered in water pH calculation (see pH section).

3.10 Bacteria

The most common bacteria met in oil and gas industry are sulfates-reducing bacteria (SRB). They
grow in anaerobic conditions by mean of sulfates ions reduction to sulfides. The microbial corrosive
attack is characterised by formation of black deposits of sulfides-containing corrosion products on
metals.

The other groups of bacteria met in oil and gas industry are:
− SOB - sulfate oxidizing bacteria: it is a group of bacteria growing in aerobic conditions and
leading to the production of sulfuric acid;
− IRB - iron related bacteria: it is a group of bacteria growing in aerobic conditions and leading to
the oxidation of iron ion from a soluble form towards one not soluble that in combination with
chlorides in the environment give place to ferric chloride that is corrosive;
− APB - acid producing bacteria: they give place to the production of organic acids and sulfuric
acid;
− ASB - aerobic slime bacteria: they give place to the formation of "slime" that is a product with a
polymeric basis.

Bacteria are present in soil, natural waters and mud; they are not normally present in hydrocarbon
reservoirs, unless by contamination occurring for instance during drilling operations or injection of
fluids into the reservoir.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 22 of 51

3.11 Sand and solid particles

The presence of sand and suspended solids in the fluid may cause erosion of metallic surface. The
erosion rate depends on fluid flow rate and density and depends on quantity, density and morphology
of the solid particles associated with the fluid.

3.12 Elemental sulfur

Elemental sulfur is sometime present in reservoir fluids. Its presence is not frequent and is usually
associated with high concentrations of H2S. It is a strong oxidant and it is very aggressive for both
carbon steel and corrosion resistant alloys.

3.13 Mercury

Mercury is sometimes detected in reservoir fluids. Like elemental sulfur, its presence is unusual.
Corrosion issues associated with mercury are mostly related to liquid metal embrittlement ‘LME’ of
susceptible alloys, typically aluminium alloys. LME caused by mercury typically affects downstream
process where aluminium alloys are widely used. Other affected materials are copper base alloys
such as Monel and gold alloys, sometimes used as brazing alloys in gas compressors. Corrosion
caused by elemental mercury on carbon and low alloy steels is not documented in the technical
literature, however corrosion from mercury compounds under specific operation conditions and in
presence of other compound such as methanol cannot be excluded and further assessment is
recommended.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 23 of 51

4. CORROSIVITY CLASSES

4.1 Hydrocarbon systems and production water

4.1.1 Corrosion parameters

The corrosivity assessment for hydrocarbon systems shall account for the following aspects:
− CO2 content;
− H2S content;
− operating and design pressure and temperature;
− oxygen content or other oxidizing agents;
− water cut and condensing conditions;
− in situ pH;
− water chemical composition, and particularly:
− chlorides content,
− total salinity,
− bicarbonates content,
− acetates/organic acids content,
− metal ion and metal concentration,
− gas oil ratio or gas liquid ratio;
− flow rates, superficial phase velocities and flow pattern;
− presence of elemental sulfur;
− oxygen contamination;
− presence of sand or suspended solids;
− presence of bacteria.

4.1.2 Corrosivity classes

Hydrocarbons systems (liquid, gas or multiphase) and production waters are classified according to
the following criteria:

non corrosive (free from CO2 and H2S) LH.N. or GH.N or PW.N
pCO2 < 0.001 bar (0. 1 kPa), and
pH2S < 0.0035 bar (0. 35 kPa).

sweet (containing CO2) LH.C. or GH.C or PW.C


pCO2 > 0.001 bar, and
pH2S < 0.0035 bar.

corrosive, sour (containing H2S and CO2) LH.CS. or GH.CS or PW.CS


pCO2 > 0.001 bar, and
pH2S > 0.0035 bar.

Dry gas is an additional attribute to gas hydrocarbon system. A gas is considered dry when the water
dew point at the actual pressure is at least 10°C lower than the actual operating temperature for the
system. Particular attention is to be dedicated to stagnant conditions. Dry gas is non corrosive,
however, a minimum corrosion allowance, for carbon steel component, and sour service
requirements are in any case recommended to account for upset or transitory conditions.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 24 of 51

4.2 Service waters

4.2.1 Corrosion parameters

The main corrosion parameters for corrosion assessment of waters are


− dissolved oxygen;
− pH;
− H2S content (only for reinjection water);
− CO2 content (only for reinjection water);
− operating/design temperature;
− total salinity aqueous phase;
− bacteria;
− hydrodynamic conditions;
− chemical analysis.

4.2.2 Corrosivity classes

Water is classified according to the following corrosivity classes:

deaerated water* W.D.


[O2] < 50 ppb (0.05 mg/l)

aerated water W.A.


[O2] > 1 ppm

Further than oxygen content, the following conditions characterize waters:


- water with low pH: 3 < pH < 6
- water with high pH: pH > 6
- H2S containing water: H2S > 1 mg/l (1 ppm)
- H2S non containing water: H2S < 1 mg/l (1 ppm)

*In conformity with Norsok M-001, temporary excursions of oxygen concentration within 200 ppb are
acceptable provided the overall fraction of time in which this condition is verified does not exceed the
10%. For poorly deaerated system, i.e. where this condition is not fulfilled, a dedicated corrosion
assessment is recommended.

4.3 Glycol

4.3.1 Corrosion parameters

Glycol corrosivity depends on the associated water and on the possibility that the associated water
will break the solution and wet the metallic surface. Affecting parameters are:
− operating/design temperature;
− hydrodynamic conditions;
− water percent;
− water data:
− dissolved oxygen;
− pH;
− H2S content;
− CO2 content;
− chlorides content (if any).
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 25 of 51

4.3.2 Corrosivity classes

Glycol can be classified as follows.

Lean (dry) glycol (water, CO2 and H2S are removed) GL.L.

Rich (wet) glycol (containing water, CO2 and H2S) GL.R.

4.4 Amine

4.4.1 Corrosion parameters

Amines are classified on the basis of the following corrosion parameters:


− CO2 and H2S content; it is expressed as acid gas loading, defined as moles of acid gas per mole
of active amine;
− amine type and concentration;
− Heat Stable Amine Salt concentration (HSAS); HSAS are amine degradation products that reduce
the amount of active amine available to absorb acid gas;
− operating/design temperature;
− operating/design pressure;
− pH;
− flow rate.

4.4.2 Corrosivity classes

Amines can be classified as follows.

Lean amine (CO2 and H2S are removed) AM.L.

Rich amine (containing CO2, H2S or both) AM.R.


Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 26 of 51

5. CORROSION FORMS

5.1 General

5.1.1 Foreword

This section summarizes the most common corrosion forms encountered in oil and gas production.

For each corrosion form, definitions and guideline criteria for corrosion prediction are provided.

5.1.2 Materials

The main metallic materials used in the oil and gas industry are classified into the following
categories:
− carbon and low alloy steels;
− stainless steels:
− martensitic;
− austenitic;
− duplex;
− nickel base alloys;
− copper alloys;
− special alloys (cobalt and titanium alloys).

Stainless steels, nickel base alloys and special alloys are usually named “CRA” which stands for
“corrosion resistant alloys”.

Non metallic materials are outside the scope of this document.

5.1.3 Corrosion morphologies

Corrosion forms can be classified into two following fundamental types, according to the morphology
of the attack.

General corrosion: it occurs on the whole surface of the metal in contact with the corrosive phase; it
can be uniform, with a generalised and regular loss of metal on the exposed surface, or non uniform,
with corrosion penetration varying from area to area (“mesa” corrosion).

The following categories are used for to express the corrosion rate of general corrosion forms:
− negligible < 50 µm/y (< 2 mpy)
− low 50÷100 µm/y (2÷5 mpy)
− moderate 100÷500 µm/y (5÷25 mpy)
− severe 500÷1000 µm/y (25÷50 mpy)
− very severe > 1000 µm/y (> 50 mpy)

Localised corrosion: corrosion preferentially concentrated at discrete sites of the surface in contact
with the corrosive environment. The morphology of localised corrosion depends on material and
environment.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 27 of 51

Environmental Assisted Cracking (EAC): it includes all kinds of stress corrosion cracking that occur
when the action of the corrosive environment is worsen by the presence of applied or residual
stresses.

5.2 CO2 corrosion

Carbon dioxide, CO2, is the main responsible for corrosion of carbon and low alloy steels in oil and
gas production. It represents the greatest risk for integrity of carbon and low alloy steel equipment in
a production environment. On the contrary, corrosion resistant alloys are generally fully resistant to
this type of corrosion.

CO2 corrosion depends primarily on the content of CO2 dissolved in the water phase. Water in
equilibrium with a CO2-containing gas phase dissolves CO2 following the Henry’s law. The CO2
partial pressure is usually known from analysis of gas separated and for this reason this parameter is
used to predict CO2 corrosion rate making reference to the proper gas/liquid equilibrium.

Dissolved in water CO2 forms carbonic acid H2CO3 which is a weak acid and dissociates only a little
into bicarbonate ion HCO3- and hydrogen ion H+. However, carbonic acid results highly corrosive for
carbon steel because the primary cathodic reaction is given directly by undissociated H2CO3 and not
by H+ as normally occurs with strong acids (H+ contributing only secondarily to cathodic control):
Primary cathodic reaction: H2CO3 + e- → H + HCO3- followed by 2H → H2
Secondary cathodic reaction: H2CO3 → H+ + HCO3- followed by H+ + e- → H2
Anodic reaction Fe → Fe2+ + 2e-
Overall reaction Fe + H2CO3 → FeCO3 (iron carbonate) + H2
Fe(HCO3)2 → FeCO3 + H2CO3
Once saturated with FeCO3 the dissolved species is not FeCO3 but Fe(HCO3)2 which at high
temperature decomposes:
Fe(HCO3)2 → FeCO3 + H2CO3
The solubility of FeCO3 is low and decreasing with temperature, but apparent solubility of FeCO3
much higher than theoretically expected was observed, due to supersaturation or slow reaction. The
iron carbonate can form a protective layer of corrosion products.

CO2 corrosion occurs with quite different morphologies, often designed with specific terms as: “mesa
corrosion”, “pitting corrosion”, “ring worm corrosion”.

CO2 corrosion is affected by for the following influencing factors:


− CO2 partial pressure;
− temperature;
− hydrodynamic conditions;
− presence of H2S;
− steel composition, and particularly chromium is beneficial;
− water phase chemical composition and “in-situ” pH.

5.2.1 Effects of chemical species in solution

Chemical composition of the water phase in contact with carbon and low alloy steels affects CO2
corrosion rate, especially by mean of the pH value modification. Corrosion rate is usually very low
when the pH value is above 5.5÷5.6.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 28 of 51

The most important effects of the chemical species are briefly reviewed here below.

Bicarbonates. In waters with high alkalinity (HCO3- = 30÷150 meq/l), corrosion rate is low or
moderate. Also the Ca2+/HCO3- ratio affects the water aggressiveness: in alkaline waters, with HCO3-
>>Ca2+, corrosion rate is low or moderate as above said; when Ca2+/HCO3- is high (> 1000
meq/meq), pH is low, but corrosion results uniform. Furthermore, when the content of Ca2+ ions in
solution is high, precipitation of a protective layer of CaCO3 can occur depending on pH and
temperature.

Acetates. Several reservoir fluids contain volatile organic acids, in particular acetic acid (CH3COOH).
With the term “acetates” it is normally intended the amount of all organic acid in solution, expressed
as meq/l. If bicarbonates are determined by titration, some acetates are titrated as bicarbonates. CO2
corrosion is affected by acetate concentration in solution. Low CO2 corrosion is expected if the
acetates content is lower than about 1 meq/l. Presence of acetates affects the equilibria that
determine the pH. Furthermore, organic acids can behave as complexing agents and under certain
conditions could lead to high corrosion rates by promoting iron dissolution and preventing the
formation of protective corrosion products.

Fe2+ ions. The presence of Fe2+ ions in water solution (i.e. from steel corrosion) increases the pH
value by means of the formation of FeCO3 and Fe3O4. The increase of the pH value is usually in the
range of 0.5 to 1.6 units when Fe2+ saturation conditions are approached. The saturation pH for
FeCO3 and Fe3O4 can be calculated as a function of temperature and CO2 partial pressure.

5.2.2 Effect of H2S on CO2 corrosion

For sour environments, compared with sweet system, H2S presence in water solution can interfere
with the mechanisms of CO2 corrosion of mild steel, mostly depending on the H2S concentration,
CO2/H2S ratio, temperature and pH.

As general guidance to the assessment of CO2-H2S systems, H2S presence on CO2 corrosion can:
− increase corrosion
o by promoting localised corrosion at a rate greater than CO2 mesa corrosion rate;
o forming FeS corrosion product that is less protective than iron carbonate
− decrease corrosion
o replacing a less protective iron carbonate film;
o forming a combined protective layer of iron sulfide and iron carbonate.

The corrosion process is governed by the dominant species.

The effect that even small concentration of H2S (10-100 ppm) can have upon CO2 corrosion is
related to complex interactions between the variety of FeS corrosion products that can forms and iron
carbonate (FeCO3). This effect might be either beneficial or not, depending on the stability of iron
sulfide scale formed.

At higher H2S concentration (above 100 ppm) and CO2/H2S lower than 20-50, as sweet corrosion is
controlled by carbonate scale layers, it might be assumed that corrosion is no longer sweet when the
corrosion product is not a carbonate but a sulfide product. The type of damage caused by H2S
appears in the form of localised corrosion or general corrosion, depending upon the type and nature
of corrosion products formed.

At present there are no generally accepted prediction algorithms for H2S corrosion.

The influence of H2S on CO2 corrosion can be summarized as follows (Ref. /4/, Ref. /5/, Ref. /6/):
− At very low level of H2S (approximately pH2S ≤ 0.001 bar), and CO2/H2S ratio higher than 200-
500, CO2 is the dominant corrosive species; at temperature above 60 °C corrosion depends on
FeCO3 formation and stability. The presence of H2S is assumed not significant on CO2 corrosion.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 29 of 51

− For hydrocarbon systems characterised by a xCO2/xH2S ratio lower than 20-50 (“H2S dominated”
systems), the formation of a metastable sulfide scale prevails on FeCO3 scale. Corrosion rates
are usually lower than in sweet system where the corrosion mechanism is controlled by CO2 and
its corrosion products. However, no algorithms are available to quantify the predicted corrosion
rate.
− For CO2/H2S ratio between 20 and 500, the contribution of H2S on CO2 corrosion, i.e. with
interactions between iron sulfide and iron carbonate, is significant. The major risk is represented
by localised corrosion due to breakdown of iron sulfide protective layer. Unless more dedicated
prediction models are available, or ad-hoc tests are carried out, the recommended approach is to
continue using conservative sweet CO2 corrosion prediction models, with the exclusion of the use
of FSCALE factor to account for mitigation effects of iron carbonate scale formation at high
temperature.

5.2.3 CO2 corrosion prediction models

Several models for CO2 corrosion rate prediction were developed and made available in the technical
literature to address the issue of durability and corrosion allowance calculation of carbon steel
components and associated risk evaluation.

Some of these models are illustrated in this section. They covers only calculation of corrosion rates
where CO2 is the corrosive agent, not including additional effects of other constituents which may
influence the corrosivity, e.g. contaminations of O2, H2S etc. If present, the effect of these
constituents is be evaluated separately.

Using the various model available, it is important note that these models were mostly based on
empirical approach, based on test data obtained varying the affecting parameters to setup the model
which best fit the observed corrosion rates, often with a significant scattering of data. For the large
number of variables influencing corrosion in the real world, their complex interaction and the
impossibility to reproduce and account for all of them during experiments.

Furthermore, the models approach is based on deterministic values of input parameters, whereas in
reality each one is associated to a level of uncertainty and the effect of some of them might be
markedly non-linear, as in the case of temperature in proximity of the scaling temperature.

For the above reasons, when using models for CO2 corrosion rate prediction it is important to bear in
mind that their applicability is only valid within a limited range of conditions and even within these
ranges, significant deviations between real corrosion rates and predicted ones cannot be excluded,
even if models attempt is usually focused on worst case scenarios.

5.2.3.1 NORSOK M-506 CO2 corrosion rate calculation model

The NORSOK M-506 model is an empirical corrosion rate model obtained correlating experimental
data for carbon steel in CO2-containing water at different temperature, pH, CO2 fugacity and wall
shear stress.

The data for the development of the model are mainly based on flow-loop tests are taken from
research programs at IFE, Institute for Energy Technology in Norway and were carried out at
temperature ranging from 5°C to 160°C. The main principles for the testing is described in a paper by
Dugstad et. al. (Ref. /12/).

The following general equations of the CO2 corrosion rate for carbon steel at each of the
temperatures (T); 5°C – 15°C – 20/40/60/80/90/120/150°C are used:

T = 5°C vNOR,5°C = KT × fCO20.62 × f(pH)T


Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 30 of 51

T = 15°C vNOR,15°C = KT × fCO20.62 × (S/19)0.146+0.0324 log(fCO2) × f(pH)T

T = 20-150°C vNOR,20-150°C = KT × fCO20.36 × (S/19)0.146+0.0324 log(fCO2) × f(pH)T

where:
− fCO2 = fugacity of CO2 (bar)
− KT = constant for temperature T
− f(pH)T = pH factor at temperature T
− S = wall shear stress (Pa)

The corrosion rate between temperature where a constant KT has been generated is found by a
linear extrapolation between the calculated corrosion rate at the temperature above and below the
desired temperature. The constant KT and the pH factor f(pH)T are given in the next table.

Temp. °C KT pH f(pH)T
3.5 < pH < 4.6 f(pH) = 2.0676 - (0.2309 × pH)
5 0.42
4.6 < pH < 6.5 f(pH) = 4.342 - (1.051 × pH) + (0.0708 × pH2)
3.5 < pH < 4.6 f(pH) = 2.0676 - (0.2309 × pH)
15 1.59
4.6 < pH < 6.5 f(pH) = 4.986 - (1.191 × pH) + (0.0708 × pH2)
3.5 < pH < 4.6 f(pH) = 2.0676 - (0.2309 × pH)
20 4.762
4.6 < pH < 6.5 f(pH) = 5.1885 - (1.2353 × pH) + (0.0708 × pH2)
3.5 < pH < 4.6 f(pH) = 2.0676 - (0.2309 × pH)
40 8.927
4.6 < pH < 6.5 f(pH) = 5.1885 - (1.2353 × pH) + (0.0708 × pH2)
3.5 < pH < 4.6 f(pH) = 1.836 - (0.1818 × pH)
60 10.695
4.6 < pH < 6.5 f(pH) = 15.444 - (6.1291 × pH) + (0.8204 × pH2) - (0.0371 × pH3)
3.5 < pH < 4.6 f(pH) = 2.6727 - (0.3636 × pH)
80 9.949
4.6 < pH < 6.5 f(pH) = 331.68 × e(-1.2618 x pH)
3.5< pH < 4.57 f(pH) = 3.1355 - (0.4673 × pH)
90 6.250 4.57< pH < 5.62 f(pH) = 21254 × e(-2.1811 x pH)
5.62 < pH < 6.5 f(pH) = 0.4014 - (0.0538 × pH)
3.5 < pH < 4.3 f(pH) = 1.5375 - (0.125 × pH)
120 7.770 4.3 < pH < 5 f(pH) = 5.9757 - (1.157 × pH)
5 < pH < 6.5 f(pH) = 0.546125 - (0.071225 × pH)
3.5 < pH < 3.8 f(pH) = 1
150 5.203 3.8 < pH < 5 f(pH) = 17.634 - (7.0945 × pH) + (0.715 × pH2)
5 < pH < 6.5 f(pH) = 0.037

If not known, the model include both a simplified and a more accurate procedure for the calculation of
the wall shear stress “S”, which is calculated as a function of liquid and gas superficial velocities,
water cut, internal diameter, density and viscosity of each phase, pipe roughness, gas compressibility
factor, the value of water cut at inversion point and the max relative liquid viscosity; in most cases
default values are provided. The model is described in details in Norsok M-506.

5.2.3.2 De Waard & Milliams empirical model (1991/1993)

The deWaard & Milliams model for CO2 corrosion prediction for mild carbon steel was originally
presented in 1975 and was further developed into a version combining the content of two papers
presented in 1991 and 1993 (Ref. /8/). The model was developed starting from a worst case
prediction with the original deWaard-Milliams equation and applying correcting factors to quantify the
influence of environmental parameters and of corrosion product scale formed under various
conditions.

Base equation. The CO2 content in the water phase in equilibrium with a CO2-containing gas phase
is proportional to its partial pressure, pCO2. The CO2 solubility in water dependents also on
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 31 of 51

temperature. Corrosion rate of CO2-containing water can be calculated with the deWaard base
equation as a function of temperature and CO2 partial pressure in the associated gas phase:

1710
log( v dWM ) = 5.8 − + 0.67 ⋅ log(pCO2)
T + 273

− vdWM corrosion rate (mm/y)


− T temperature (°C)
− pCO2 CO2 partial pressure (bar)

The corrosion rate predicted by the base equation is often an overestimation of reality. To take into
account the effect of some influencing parameters, the “base” corrosion rate calculated with the base
formula have to be adjusted by relevant corrective factors.

Effect of total pressure. At high pressure, the non-ideality of gas shall be considered, and the CO2
fugacity, fCO2, is to be used instead of pCO2 in the base formula. The correction introduced becomes
appreciable only when pressure is above 100 bar.

Effect of temperature on scaling formation. Above a certain temperature, named scaling temperature
(TSCALE), the corrosion rate decreases by virtue of the formation of stable corrosion products with
protective characteristics, unless hydrodynamic conditions are not so severe to cause their removal
by erosive action. To account for this effect, a corrective factor FSCALE ≤ 1 is introduced in corrosion
rate calculation when the scale temperature, TSCALE, is exceeded and the scale factor FSCALE are
given by the following equations:
2400
TSCALE [°C] = − 273
6.7 + 0.60 ⋅ log( fCO 2)
2400
− 0.60 ⋅log( fCO2 ) − 6.7
FSCALE = 10 T + 273

Even if at temperature exceeding the TSCALE the corrosion rate tends to decrease to zero with time,
the reliability of a complete protection afforded by scale cannot be sufficiently quantified and the
scale factor gives a minimum estimate for its protectiveness. Furthermore, the scale layer might be
weakened in high chloride concentration solutions, by the presence of organic acids or can be
eroded in high flow rate regimes. For these reasons it is suggested to set FSCALE = 1 in case
formation water is present, in presence of organic acids, when superficial gas velocity in wet gas
systems exceeds 20 m/s or in multiphase systems at the onset of annular mist flow regime.

Effect of pH. At a temperature lower than TSCALE and in absence of other buffering chemical species,
the anodic reaction delivers in solution Fe2+ ions and consumes H+ ions until the solution is saturated
with Fe2CO3 or Fe3O4. This precipitation can lead to the formation of a protective layer or not. When
Fe2CO3 saturation occurs, further addition of iron will not change the Fe2+ concentration nor solution
pH (pHSAT). In case of saturation with Fe3O4 Fe2+ keep increasing. The contamination of the CO2
solution with corrosion products leads to a pH shift and to a reduction of corrosion rate. The influence
of pH shift on corrosion rate is taken into accounted by means of a corrective factor “FpH” calculated
as follows.

The initial pH of unbuffered solution of water and CO2 only is given by:
pH(water+CO2 only) = 3.71 + 0.00417×T – 0.5log(fCO2)
The saturation pH, pHsat, is calculated for the precipitation of FeCO3 and Fe3O4:
1307
pHsat,FeCO3 = 5.4 − 0.66 ⋅ log(fCO 2 ) or pH sat,Fe3O 4 = 1.36 + − 0.17 ⋅ log( fCO 2 )
T + 273
pHSAT = minimum value between (pHsat,Fe3CO3, pHsat,Fe3O4)
pHSAT – pH(water+CO2 only) = 0.5 ÷ 1.6
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 32 of 51

The corrosion product with the lowest saturation pH is the most stable thus the most likely to form.
The corrective factor, FpH, is given by:

if (pHsat – pHactual) > 0: FpH = 10 0.32 ⋅(pHSAT − pHACTUAL )


1.6
if (pHsat – pHactual) < 0: FpH = 10 −0.13 ⋅(pHACTUAL − pHSAT )

Effect of liquid hydrocarbon. Effect of crude oil presence, with the capability to keep water entrained
into oil emulsion avoiding water wetting of the metal surface is taken into account by a oil factor FOIL,
which assume the value FOIL = 0 if the water cut is lower than 30% and the if the oil flowrate is above
1 m/s. If both these conditions are not verified FOIL = 1 shall be set. Liquid hydrocarbons condensates
are poorly capable to entrap water into oil emulsion, irrespectively of the water content.

Condensation Factor. Corrosion rate of steel exposed to condensing water phase quickly decreases
with time. This is due to the fact that liquid water film formed by gas condensation in slow cooling
systems are quickly saturated with corrosion products. For wet gas systems cooling slowly, as
normally occur on gas transport systems, a condensation factor FCOND = 0.1 can be used to correct
base corrosion rate to account of this fact. For fast cooling systems, as in the case of heat exchanger
tubes, condensation rate is too high (above 0.25 g/m2s) to give place to this effect and FCOND=1 shall
be set.

5.2.3.3 De Waard, Lotz, & Dugstadt semi-empirical model (1995)

CO2 corrosion rate of carbon steel observed from a large number of experiments in a high pressure
test loop with various strictly controlled environments and flow conditions were fitted to a semi-
empirical model equation (Ref. /10/). This model uses a resistance model to combine a) the
contribution of flow independent kinetics of the corrosion reaction and b) the contribution of flow
dependent mass transfer of dissolved CO2 influenced by fluid velocity.

Limits of model validity. The tests included measurements within a maximum temperature of 90°C
and maximum pCO2 partial pressure of 20 bar. This may limit the model validation for high
temperature and high pressure applications.

Base equation. CO2 corrosion rate is calculated by the following procedure:


1 1 1
= +
v dWLD vR vM

where vR and vM are calculated as follows:


1119
log( v R ) = 6.23 − + 0.0013 ⋅ T + 0.41 ⋅ log(pCO 2 ) − 0.34 ⋅ pH ACT
T + 273
U0.8
v M = 2.45 ⋅ 0.2 ⋅ pCO 2
D
− vR reaction corrosion rate (mm/y);
− vM mass transfer corrosion rate (mm/y);
− U liquid phase velocity (m/s). If the liquid phase velocity is unknown, the average fluid velocity is used;
− D hydraulic diameter (m). If the hydraulic diameter is unknown, the internal diameter is used.

The base equation incorporates pH effects, but it is to be modified by applying a correction factor to
account for the effects of protective corrosion products formation.

Effect of temperature on scaling (corrosion products) formation. As previously seen the formation of
protective corrosion products has to be considered above a certain temperature. The correction
factor is applied when T > TSCALE:
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 33 of 51

2400
TSCALE = − 273
6.7 + 0.44 ⋅ log( pCO 2)
and it is calculated through the following formulae:
2400
log(FSCALE ) = − 0.44 ⋅ log(pCO 2) − 6.7
T + 273

The FSCALE factor is applied to the term vR only.

Corrected equation. The potential corrosion rate is then calculated multiplying the base corrosion rate
by the correction factor FSCALE:
v × v ⋅F
v dWLD,C = M R SCALE
v M + v R ⋅ FSCALE

5.2.3.4 Effects of corrosion inhibition treatment and glycol

Effect of glycol. Glycol affects CO2 corrosion in two ways:


− by reducing the potential corrosivity of the water phase (increasing pH etc.);
− by absorbing water from the gas phase.

The effect of glycol on corrosion rate is taken into account by means of a glycol factor FGLY, ranging
from 0.008 to 1, and calculated with the following formula:

FGLY = 101.6 × (log (100-Gly%) – 2)

− Gly% is the weight percentage of glycol;


− A is a constant which depend only weakly on type of glycol and is normally assumed A=1.6 for all types.

For glycol weight content above 95%, FGLY is set to 0.008 (see Norsok M-506).

The corrosion reduction factor FGLY shall not be used in combination with corrosion inhibitor efficiency
for corrosion prediction purposes (see “corrosion rate correction” at the end of this paragraph).

Effect of corrosion inhibitor. If corrosion inhibitor is injected into the system, corrosion rate evaluation
shall take into account of inhibitor efficiency and inhibitor availability.

Corrosion inhibitor availability “A%” is defined as the time the inhibitor is present in the system at a
concentration at or above the minimum dosage. The percentage availability (A%) is defined as:

inhibitor _ available _ time


A % = 100 ×
design _ life _ time

The inhibitor availability depends on the corrosion management program, including corrosion
monitoring strategy. An inhibitor availability of 90% is recommended. A maximum value of 95% can
be considered if supported by consistent corrosion management system including qualified inhibitor
injection and corrosion monitoring in place since production startup to actively monitor corrosion and
inhibitor injection.

The residual corrosion rate CRINHIB during proper treatment should be then considered. At the design
stage, before data from corrosion inhibitor testing are made available, it may be assumed that
inhibition can decrease the corrosion rate to 0.1 mm/year. The inhibited corrosion rate shall,
however, be documented by corrosion tests at the actual conditions or by field test data.

The overall corrosion inhibition treatment efficiency CIE% over the entire lifespan is given by the
following expression:
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 34 of 51

CIE% = A% × (1 − CRINHIB / CR UNINHIB )

Use of corrosion inhibitors in process systems can be used provided the inhibitor in each process
stream satisfies the inhibitor supplier's minimum recommended concentration for each stream and
flow rate. Due to complex geometries and normally high flow rates, there is an increased risk for high
inhibited corrosion rates locally in process systems compared to pipelines, which will influence the
need for inspection and maintenance.

Corrosion rate correction. When both corrosion inhibitors and glycol are injected, only the treatment
giving the greatest corrosion rate reduction factor will be used for calculation of the resulting
corrosion rate.

5.2.4 Top of line corrosion

In horizontal flow, in the space above a stratified liquid, water condensation from the gas phase may
occur, causing corrosion localised at the top of the line (TOL) which, especially if corrosion inhibition
treatment is performed, may be more severe than at the bottom because inhibitor is not effective on
metal surface wetted by condensing water. For slow cooling system, a condensation factor FCOND =
0.1 as described in paragraph 5.2.3.1 is applicable.

5.2.5 Corrosion products

The prediction of characteristics of corrosion products forming on steel in CO2-containing fluids is


useful to predict the morphology and the severity of corrosion attack. As seen in paragraph 5.2.2, the
nature of corrosion products can differ significantly between sweet systems and sour systems and
particularly on the basis of the CO2/H2S ratio.

Sweet systems (CO2/H2S ratio > 200-500)


In sweet system or in systems with a CO2/H2S ratio above 500, the corrosion products consists of
iron carbonates as illustrated paragraph 5.2.2. As a rule of thumb the following criteria can be
applied:
− T < 60 °C and pCO2 < 5 bar (0.5 MPa): formation of amorphous FeCO3 with poor adherence;
− 60 °C < T < 100 °C: formation of a protective layer of FeCO3;
− T > 100 °C: the layer becomes more stable and increases considerably its protectiveness and
formation of Fe3O4 occurs.

The protective effect of corrosion products is also related to hydrodynamic conditions. High flow rates
with associated high shear stress on the metal surface, could lead to a remove of the protective
corrosion products by erosive wear and a consequent increase in corrosion rate.

Sour systems (CO2/H2S ratio < 20-50)


Sour systems characterised by a CO2/H2S ratio lower than 20 they consists of iron sulfides. The
following reactions of steel with wet H2S are generally proposed:
Fe + H2S + H2O = (FeHS-)ads + H3O+
FeHS-ads → FeHS+ads + 2e-
FeHS+ads → FeS1-X + xSH- + (1-x)H+

Where the product FeS1-X is predominantly mackinawite, Fe9S8 (tetragonal FeS), or pyrite at lower
pH, from 20 to 90°C. At temperature higher than 100°C very protective pyrrhotite forms. The nature
of sulfide scales is however very difficult to predict, because of the complexity of the various sulfides
that can form depending on pH and temperature.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 35 of 51

Iron sulfide scales stability appears to be strongly dependent on chlorides or elemental sulfur
presence and more sensitive than iron carbonate scales to breakdown under turbulent conditions. A
critical chloride threshold in the range 10,000-50,000 ppm, above which chloride ions can destroy the
protective FeS scale leading to an increased corrosion rate has been proposed based on field and
laboratory data (Ref. /7/).

A key factor is also played by oxygen in case of fluid contamination, by changing the corrosion
products and significantly affecting the resulting corrosion rate. Oxygen contamination in wet H2S
may determine quasi-uniform corrosion rates in the order of 0.5 to 1 mm/y or even more serious
risks, by formation of thiosulfates or, in presence of chlorides, promoting localised corrosion at higher
corrosion rates.

Intermediate systems (200-500 > CO2/H2S ratio > 20-50)


Between 500 and 20, corrosion products consists of a combination of iron carbonates and iron
sulfides, displaying an effect on corrosion rates that is difficult to predict.

Protective layers are always thin as they progressively reduce ionic transport and corrosion reaction,
whereas non-protective layer are normally thick and abundant. Corrosion product layer thickness
could be an indicator of its protectiveness.

5.3 Sour Service

Sour service is associate with the presence of H2S in the oilfield environment. Exposure of materials
to sour environments can cause cracking mechanism on susceptible materials. These mechanisms
include sulfide stress cracking “SSC”, stress corrosion cracking “SCC”, hydrogen-induced cracking
“HIC” or stepwise cracking “SWC.

In H2S containing fluid, sour service resistant materials shall be used with material requirements and
environmental application limits in accordance with ISO 15156 / NACE MR0175.

5.3.1 Sulfide Stress Cracking (SSC)

Sulfide Stress Cracking (SSC) occurs on susceptible materials when specific environmental
conditions are met, characterised by the presence of H2S, water and mechanical stresses, applied or
residual.

SSC is a form of hydrogen stress cracking resulting from adsorption of atomic hydrogen, produced
by the corrosion reaction on the metal surface in presence of H2S. The atomic hydrogen can diffuse
into the metal but remains in solid solution in the crystal lattice. The material ductility is impaired and
the susceptibility to cracking is increased.

SSC is influenced by a complex interaction of affecting parameters, including:


− chemical composition, strength, heat treatment and microstructure of the material;
− pH;
− hydrogen sulfide concentration and total pressure;
− total tensile stress (applied plus residual);
− water wetting conditions;
− temperature;
− time.

SSC usually occurs more readily in high-strength and high-hardness steels, in hard weld deposits or
in hard heat affected zones of lower-strength steels. Post weld heat treatment at about 620°C or
higher are often specified to removed residual stress induced by welding.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 36 of 51

The severity of the sour environment with respect to SSC of carbon or low alloy steels is assessed in
accordance with ISO 15156 by using Figure 5.1.

7.5
0.0035 bar
(0.05 psi)

SSC Region 1
6.5 (Slightly Sour Service)

SSC Region 2
(Moderately Sour Service)
5.5 SSC Region 0
(Sweet Service)
In-situ pH

4.5

SSC Region 3
(Severely Sour Service)

3.5

2.5
0.0001 0.001 0.01 0.1 1 10 100
Hydrogen Sulphide Partial Pressure (pH2S - bar)

Figure 5.1 - Sour service conditions according to ISO 15156 / NACE MR0175

5.3.2 Sour Service Resistant Materials

In sour service, materials resistant to SSC (sour service grades) shall be specified with the criteria for
selection and qualification provided by ISO 15156 / NACE MR0175 and Company Specification
03587.MAT.COR.PRG “Metallic materials in contact with H2S containing environments. Corrosion
tests methods and evaluation criteria”.

Material requirements for carbon and low alloy steel pressure vessels in sour service shall be in
accordance with Company Specification 05489.MAT.COR.SDS “Additional Requirements for
Pressure Vessels for Applications in H2S Containing Environments”.

Sour service material requirements and environmental application limit for corrosion resistant alloys,
CRAs, are provided by ISO 15156-3 and are based on the following affecting parameters:
Environmental parameters:
− H2S partial pressure;
− water phase pH;
− temperature;
− water phase chloride concentration;
− presence of elemental sulfur in the environment;
Material parameters:
− method of manufacture and finished condition of the material;
− pitting resistance of the material;
− welding and post weld heat treatments;
− heat treatments;
− hardness and mechanical strength;
− etc.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 37 of 51

5.3.3 Hydrogen Induced Cracking (HIC)

Hydrogen Induced Cracking (HIC) it is intended the deterioration of the properties of the material
caused by atomic hydrogen that enters the metallic material and which may arise without any
external applied or residual stress. Metal damages can occur as surface blistering or as formation of
internal stepwise microcracks.

The terms used to define such cracking include:


− blistering;
− internal cracking;
− stress oriented hydrogen induced cracking;
− stepwise cracking (SWC);
− hydrogen induced cracking (HIC);
− hydrogen pressure induced cracking (HPIC).

Formation of microcracks or blisters is caused by atomic hydrogen produced on the corroding metal
surface which, in presence of H2S, diffuses into the metal lattice. The atomic hydrogen collects at
elongated inclusions or segregation bands of the microstructure and recombines to molecular
hydrogen, producing internal stresses leading to the mentioned metal damages.

Control of microstructure and particularly the cleanliness of steels reduce the availability of crack
initiation sites and are therefore critical to the control of this phenomenon.

Materials susceptible to HIC are typically carbon and low alloy steels produced by lamination and
containing S and Mn in sufficient amounts to give place to inclusions, and particularly C-Mn steels
with manganese sulfide inclusions, MnS type II.

The probability of HIC is influenced by steel chemistry and manufacturing route. HIC is controlled by
limiting the level of sulfur in alloy.

Flat rolled steel. Flat rolled steel displaying adequate resistance to HIC are mainly obtained by
limiting the content of sulfur below 30 ppm of ([S]max = 30 ppm). Calcium treatment (Ca/S > 2) or rare
earth treatment producing spheroidal inclusions are also beneficial.

Seamless pipes and forgings. Seamless pipe are much less sensitive to HIC. For conventional, hot
rolled, seamless products the following limitation of sulfur content in the alloy is considered sufficient
to avoid HIC:
− seamless pipes: [S]max = 100 ppm;
− forgings, fittings: [S]max = 250 ppm.

Casting. Casting product with a sulfur content lower than 250 ppm ([S]max = 250 ppm) are normally
not sensitive to HIC.

5.3.4 H2S corrosion

The presence of H2S in the fluid affects CO2 corrosion as discussed in paragraph 5.2.2.

5.4 Pitting and crevice

Pitting and crevice are localised corrosion forms typical of metals showing active-passive behaviour.
Certain circumstances could promote a local breakdown of the passive layer which leads to the set-
up of a galvanic macro-cell between the depassivated region of the metal where anodic dissolution
occurs and the surrounding region, which remain passive and where the cathodic process takes
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 38 of 51

place. The anodic area is always very small and the metal dissolution is enhanced by local
acidification and by the unfavourable ratio of the coupled galvanic areas, which lead to very high
penetration rates.

In pitting corrosion, the anodic area is the bottom of the pit, while in crevice corrosion it is the
shielded metal surface within the crevice.

The susceptible materials are those ones that normally operate in passive conditions, and
particularly:
− stainless steels;
− nickel alloys;
− copper alloys.

To provide a comparative indication of the pitting resistance specific parameters based on alloy
composition have been proposed. These indexes are not valid for all materials, but only for austenitic
and duplex stainless steels and for some nickel base alloys.

The Pitting Resistance Equivalent (PRE) is calculated from the chemical composition as follows:

PRE = Cr% + 3.3 (Mo% + 0.5 W%) + 16 N%

For crevice corrosion, the Critical Crevice Index (CCI) is used:

CCI = Cr% + 4.1 Mo% + 27 N%

The pitting and crevice indexes are often used as absolute values, but the most appropriate use is as
ranking parameters of different materials. A typical reference index of resistance to pitting in
seawater at ambient temperature is given by a PRE > 40, which is the required value for a
superduplex stainless steel.

The Critical Pitting Temperature (CPT) and the Critical Crevice Temperature (CCT) are indexes
experimentally determined by immersion tests in FeCl3 (see ASTM G48-76) or by electrochemical
methods. Good correlation between CPT vs. PRE and CCT vs. CCI were found for austenitic and
duplex stainless steels.

5.4.1 Initiation and propagation conditions

Initiation of pitting and crevice corrosion is typically promoted by the presence of chloride ions in
solution. For AISI 300 series austenitic stainless steels (as types and 316) the safe threshold for
chloride concentration to avoid pitting and crevice corrosion in aerated aqueous solution is within 50
ppm (AISI 304) and 200 ppm (AISI 316). Above these values, the suitability of these types of
materials has to be evaluated with respect to temperature, oxygen presence, flowing conditions etc.

Availability of a cathodic process, i.e. oxygen reduction, hydrogen evolution, elemental sulfur
reduction, is needed for propagation to occur.

Hydrodynamic conditions have a great influence on pitting initiation: AISI 316 stainless steel, for
instance, is resistant to pitting corrosion in sea water provided that flow rate is above 1.5 m/s, while in
stagnating conditions pitting corrosion occurs.

Tendency to pitting and crevice greatly increases above ambient temperature.


Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 39 of 51

5.5 Stress Corrosion Cracking (SCC)

Further than sulfide stress cracking in sour service, other forms of stress corrosion cracking can
occur in oil and gas production and service facilities:
− SCC of stainless steels in chloride containing environments;
− SCC in presence of polythionic acids;
− SCC in presence of amine solutions.

5.5.1 Chloride Stress Corrosion Cracking

Chloride stress corrosion cracking is a main concern for AISI 300 series austenitic stainless steels
(i.e. AISI 304 and 316), duplex stainless steels are also susceptible, while nickel alloys are practically
immune, especially when oxygen contamination is excluded.

Austenitic stainless steels. SCC in austenitic stainless steels type AISI 304 and 316 is observed in
the following conditions:
− T > 60°C;
− Cl- > 10 ppm;
− applied or residual stress state, σ > 150 MPa (σ > 30% σYS).

With respect to the most influencing parameters, temperature and chloride concentration, the
following chlorides concentration limits can be assumed for stainless steels type AISI 304 or 316:
− T < 50 °C Cl- with no limits
− 50 °C < T < 100 °C Cl- < 100 ppm
− 100 °C < T < 150 °C Cl- < 30 ppm
− T > 150 °C Cl- < 10 ppm

The indicated chlorides concentration limits have to be used carefully, because of possible local
concentration phenomena. In practical terms, it is recommended to restrict the use of austenitic
stainless steels within a maximum operating temperature of 50 °C.

High alloy austenitic stainless steels, characterised by a nickel content above 15%, display a higher
resistance to SCC.

Other influencing parameters are:


− dissolved oxygen in solution promotes SCC;
− the lower the pH the lower the time to failure (at pH lower than 2 general corrosion prevails and
SCC does not occur);
− the higher the tensile stress the lower the time to failure;
− cold-working beyond a certain degree (around 20%) might increase significantly susceptibility to
SCC;
− alloying elements: increase of nickel above 10% and silicon above 2% are usually beneficial;
harmful elements are phosphorus and nitrogen, which have a synergistic negative effect; their
content should be limited as follows: P < 0.005% and N < 0.02%;
− welds are preferential sites for SCC initiation because of residual stress conditions and because
of their metallurgical criticality.

Ferritic and martensitic stainless steels. Stainless steels with ferritic or martensitic microstructure are
less susceptible to stress corrosion, although not immune, because of the low nickel content (Ni <
1%). The low nickel content, in fact, leads to a lower stability of the passive film and promotes
uniform corrosion.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 40 of 51

Duplex stainless steels. Biphasic austenite-ferrite stainless steels display higher resistance to SCC,
compared to ferritic and austenitic types. The optimum resistance conditions are obtained when the
two phases are present in balanced quantities and for the highest nickel contents.

Nickel based alloys. Nickel alloys are pretty resistant to chloride stress corrosion; alloys with nickel
contents above 45% are practically immune.

5.5.2 Polythionic Acid Stress Corrosion Cracking

During periodical plant shutdown, oxygen contamination by entrance of air and humidity may occur,
and polythionic acids (H2SxO6 with x = 3, 4 or 5) can form by oxidation of iron sulfide:
8FeS + 11O2 + 2H2O = 4Fe2O3 + 2H2S4O6
Polythionic acids stress corrosion is more frequent in refinery, particularly in desulfuration units.

Austenitic stainless steels, especially if sensitised, are quite susceptible. Failure occurs by crack
formation, predominantely intergranular but transgranular cracks may occur if chlorides are present.

Duplex stainless steels display a higher resistance than series AISI 300 austenitic stainless steels.
Nickel alloys (800 and 600 series) are not reported to be sensitive.

Prevention is based on operation control during plant shut downs, e.g. by means of pure nitrogen
blanketing to prevent oxygen entrance.

Recommended materials are stabilised stainless steels (AISI 321 and 347). Heat treatment (900°C
for at least 20 minutes) of weld for stabilisation is also recommended.

5.5.3 Amine (or Alkaline) Stress Corrosion Cracking

Amine Stress Corrosion Cracking, also referred to as Alkaline Stress Corrosion Cracking is a form of
stress corrosion craking occurring on low alloy and carbon steels in presence of an aqueous
alkanolamine solution at high temperature. It is typically observed in amine treating units, where
aqueous alkanolamine solutions are used to remove H2S and CO2 from hydrocarbons.

The cracking is predominately intergranular, and typically occurs in carbon and low alloy steels as a
network of very fine cracks filled with corrosion product filled. Cracking is mainly associated with the
weld heat affected zones (HAZ).

Amine cracking occurs with a variety of steels and there is no significant correlation between material
susceptibility and material properties. For instance, the steel hardness has virtually no effect.

Four available parameters are used to assess the susceptibility of steels fabrications to amine
cracking:
− type of amine;
− amine solution composition;
− metal temperature;
− tensile stress level.

Type of amine. With regard to the type of amine, results of a NACE survey indicate that amine
cracking is most likely in monoethanolamine (MEA) and disopropanolamine (DIPA) units and to a
lesser extent in diethanolamine (DEA) units. Cracking is less frequent in methildiethanolamine
(MDEA) and diglycolamine (DGA) units.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 41 of 51

Amine solution composition. Cracking occurs in a narrow range of electrochemical potential which is
very dependent upon the amine solution composition. Contaminants such as carbonate, chlorides
and cyanides have been shown to affect cracking susceptibility. Cracking typically occurs in the lean
alkanolamine solution, with a pH in the 8 to 11 range and that contains very low level of acid gases.
Amine cracking does not occur in fresh amine solutions, i.e., those that have not been exposed to
acid gases. In rich amine solutions, other forms of cracking are far more prevalent (SSC, HIC).

Metal temperature. Amine cracking can occur over a wide range of temperatures, but susceptibility is
generally higher as temperature increases. A key consideration is the actual metal temperature, and
not just the normal process temperature. Cracking has occurred in equipment and piping that
normally operates at low temperatures but was heat traced or streamed out prior to water washing to
remove residual amine solution.

Level of tensile stress. As-welded or low alloy steel fabrications are susceptible to amine cracking
because of the high level of residual stress remaining after fabrication by these methods. Application
of a post fabrication stress relieving heat treatment, for example postweld heat treatment (PWHT), is
a proven method of preventing amine cracking. For MEA and DIPA units PWHT is recommended for
all carbon steel equipment, including piping, exposed to amine solutions regardless of service
temperature. For DEA units PWHT is recommended for service temperatures higher than 60°C and
for MDEA units for service temperatures higher than 82°C.

5.6 Oxygen corrosion

Oxygen corrosion depends on:


a) environmental conditions:
− oxygen concentration;
− hydrodynamic conditions;
− temperature.
b) steel surface conditions.

Oxygen corrosion is expected to occur in the following conditions:


− aerated water;
− poorly deaerated water, due to residual dissolved oxygen.

5.6.1 Carbon and low alloy steels. Aerated water

Oxygen corrosion of carbon and low alloy steels occurs in aerated environments as general
corrosion. Corrosion rate is proportional to the amount of oxygen available at the metal solution
interface, i.e. the oxygen flux. Corrosion rate is generally mitigated by the formation of partially
protective layers of corrosion products.

If chlorine is present, its contribution to corrosion is similar and stronger. The amount of oxygen and
chlorine dissolved in water is expressed by the oxygen equivalent, [O2]EQ, defined as:
[O 2 ] EQ = [O 2 ] + 0.3 × [Cl 2 ]

The amount of oxygen dissolved in water depends on temperature, O2 partial pressure in the gas
phase and salinity.

Oxygen solubility:
− decreases with temperature;
− increases with pressure;
− decreases with salinity.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 42 of 51

The average oxygen concentration for natural aerated water is 5÷10 ppm.

The concentration of O2 dissolved in water should be periodically measured by water analysis.

In stagnant condition (flow velocity lower than 1 m/s) and at room temperature, corrosion rate,
expressed in µm/y, is about 20 times the oxygen concentration expressed in ppm.

In flowing conditions (above 1 m/s) and at different temperature, corrosion rate varies in accordance
to the following formula:
vcorr = 20 ⋅ [O2]EQ ⋅ 2(T-30)/30 ⋅ Un
where:
− vcorr corrosion rate, (µm/y);
− [O2]EQ oxygen equivalent concentration, (ppm);
− U flow rate, (m/s) (it is assumed to be U=1 m/s if U<1 m/s);
− T temperature, (°C) (it is assumed to be T=30 °C if T<30 °C);
− n 0.5 for laminar flow and 1 for turbulent flow.

This empirical equation, based on the Fick’s law, accounts for the oxygen diffusion in liquid phase
and the kinetic effects with temperature. The corrosion rate returned by this formula is reliable for
oxygen concentration above 3 ppm. Below this value the formula underestimates the actual oxygen
corrosion rate.

5.6.2 Carbon and low alloy steels. Deaerated water

With deaerated water it is intended originally aerated water where oxygen was removed by physical
or chemical methods, or by a combination of them.

In conformity with requirements specified in Norsok M-001 oxygen content in deaerated water shall
be less than 50 ppb. Temporary excursions up to 200 ppb are admitted for limited periods (less than
10% operative time over annual basis). In such conditions, deaerated water is in practical terms not
corrosive.

Oxygen corrosion from deaerated water is caused by poor deaeration resulting in an excess of
residual oxygen content. If the oxygen content exceeds 200 ppb or 50 ppb for extended periods of
time, water has to be considered poorly deaerated and some corrosion has to be expected. General
corrosion rates in the order of 0.1 mm/y and the risk of localised corrosion, galvanic corrosion and
stress corrosion cracking shall be accounted for.

5.6.3 Corrosion resistant alloys

The effect of oxygen presence (or oxygen contamination) on CRAs is associated to the risk of
localized corrosion and stress corrosion cracking. Particular care shall be considered in presence of
sour service fluid, with respect to the risk of elemental sulfur formation by H2S oxidation.

Austenitic stainless steels. Austenitic stainless steels are susceptible to localized pitting corrosion in
the presence of small amounts of oxygen. Furthermore, these materials are extremely susceptible to
chloride stress corrosion cracking and cannot be used at temperature above 60 °C under applied or
residual stress.

Duplex stainless steels. Duplex stainless steels are susceptible to localized pitting, crevice corrosion
and stress corrosion cracking in presence of oxygen and chlorides. Types 22 Cr and 25 Cr are not
recommended in aerated water.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 43 of 51

Martensitic stainless steels and precipitated hardened stainless steels. These types of alloys display
a low corrosion resistance in aerated water where they are susceptible to localized crevice and
pitting corrosion.

Nickel based alloys. In aerated systems or in poorly deaerated systems, localised corrosion and
stress corrosion cracking of nickel based alloys has to be assessed.

5.7 Erosion-corrosion and erosive wear

Erosion-corrosion is the result of the interaction of electrochemical corrosion and mechanical wear of
the flowing fluid on the metal surface, where high flow rates and high local turbulence could lead to a
local breakdown or a complete removal of the protective corrosion products from the metal surfaces,
exposing the bare metal to the corrosive environment.

The term “erosion” is pertinent to the removal of material from a metal surface, due to any
phenomenon correlated with fluid flow. The term “flow accelerated corrosion” applies when the flow
of a corrosive fluid has the only effect of accelerating the electrochemical corrosion processes
without a mechanical damage of the metal surface.

The above distinctions, introduces the following classifications:


− Corrosive / Non-Corrosive systems;
− Totally Solid-Free / Nominally Solid-Free / Solid Particles (Sand) Laden Systems;
− Multiphase / Liquid / Gas systems.

The combination of the above categories lead to different scenarios from the point of view of flow
related aspects.

5.7.1 Totally solid-free systems

The definition “totally solid-free systems” applies when the presence of solid particles transported
with the fluid can be completely excluded. In solid-free systems flow velocity can lead to the set-up of
pure abrasion or erosion-corrosion mechanisms under certain circumstances.

For multiphase (or wet gas) non-corrosive or corrosion-inhibited systems, abrasion conditions may
occur when liquid droplets impingement reach a certain level of energy. This condition is associated
with superficial gas velocity in excess of 70 m/s in a disturbed or annular flow pattern. For corrosion-
inhibited system the average flow velocity shall be limited to allow inhibitor to work properly. The
proposed limit (UMAX) is the minimum between 20 m/s and 200/√ρm.

For multiphase (or wet gas) corrosive systems, erosion-corrosion conditions is again related to liquid
droplets impingement. This condition is associated with a maximum allowable flow velocity. The
maximum erosion-corrosion flow velocity UMAX proposed for nominally solid-free systems are
recommended.

For monophase liquid systems or dry gas system, erosion-corrosion is not expected and no velocity
limitations are deemed necessary.

5.7.2 Nominally solid-free systems

The definition “nominally solid-free systems” applies when the solid particles mass flow rate in the
fluid is lower than the detectable threshold (which is about 0.5 kg per 1000 barrels in
liquid/multiphase systems and 1.5 kg per million of std.m3 in gas systems) cannot be excluded
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 44 of 51

For corrosive fluids, erosion-corrosion phenomena gives place to characteristics attacks presenting
different morphologies as shallow pits, horseshoes, or other local phenomena correlated with the
direction of flow. Erosion-corrosion initiates if the average fluid velocity in wet gas or multiphase
systems is above a critical value given by the following formula:

C
UMAX =
ρm
where:
− UMAX = erosion-corrosion critical velocity;
− C = erosion-corrosion constant;
3
− ρm = fluid density at operating conditions [lb/ft ].

The erosion constant C can assume different values depending on material and corrosion inhibitor
injection. In the Metric system, the units for the C constant are (kg/ms2)1/2, while in Imperial units are
(lb/ft.s2)1/2, and the conversion factor from Imperial to Metric is 1.22.

The following values for the C constant, expressed here in Imperial and in Metric (in brakets) units,
are recommended:
− C = 135 (165) carbon steel, not-inhibited;
− C = 200 (244) carbon steel, inhibited;
− C = 300 (366) martensitic stainless steel;
− C = 350 (427) duplex stainless steel;
− C = 488 (400) highly alloyed superaustenitics or nickel base alloys.

5.7.3 Solid particles laden systems

In presence of solid particles, the erosion corrosion is accelerated by the mechanical interaction
between the metal surface and the solid particles dragged with the fluid.

The erosivity threshold above which the scale is damaged is defined erosion-corrosion resistance.
The erosion-corrosion resistance depends on environmental factors (such as pH, temperature, CO2
partial pressure, flow velocity, solid particles mass concentration and size) and on pipe geometry. For
corrosion resistant alloys, the higher the protective film regeneration rate, the higher the threshold
erosivity. The use of corrosion inhibitors can increase the threshold erosivity.

When erosivity is high enough, the resulting erosion-corrosion wastage rate WR can be analysed by
breaking down the phenomenon into the two contributions, i.e. erosion rate ER and corrosion rate
CR as follows:
WR = ER + CR
ER can be calculated by using specific sand corrosion prediction models. Three different scenario
can be recognised, based on the calculated value of the erosion rate severity ER, which determine
the method to calculate the corrosion rate contribution (Ref. /16/, Ref. /17/, Ref. /18/, Ref. /19/, Ref.
/20/).

Scaling regime. For low erosive systems, indicatively ER < 0.1 mm/y, the protective corrosion
product layer is not damaged by the impinging particles and keeps adherent to the metallic surface,
maintaining a certain degree of protection from corrosion. The erosion rate is negligible or is low
enough to allow the scale to regenerate and the corrosion rate is limited by scaling. In this case there
is not synergistic effect between erosion and corrosion. WR is given by the sum of ER, and CR which
is calculated with the applicable corrosion prediction model.
WR = ERSCALING + vCR,SCALE
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 45 of 51

General wastage regime. For highly erosive systems, where the impinging particles remove the
protective corrosion scale layer and prevent it from regenerating or where the corrosion product are
not stable. The erosion rate is indicatively above 1 mm/y and the corrosion rate is that one of fresh
bare metal. In this case there is a synergistic effect between erosion and corrosion, which is given by
the corrosion rate of the bare metal with no protection from corrosion products and with the flow
accelerated effects related to mass transport enhancement. The corrosion rate calculation should
account for the effect of flow velocity in mass transfer on corrosion reaction and should not include
corrosion reduction factors related to corrosion products formation.
WR = ER + vCR,BARE
Pitting regime. For intermediate erosivities, the impinging particles cause the local breakdown of
corrosion product layer. The bare metal zones act as anodic areas whereas the scale-covered zones
act as cathodic areas and pitting corrosion can occur and very high penetration rates can result (in
the order of 10 mm/year). It may be assumed that in these conditions the corrosion rate contribution
is two times higher than general wastage regime, and the overall erosion-corrosion rate can be
calculated as follows:
WR = ER + 2 × vCR,BARE

Erosive wear ER due solid presence is related to the following parameters:


− content of solid particles in the fluid;
− flow rate;
− fluid density;
− fluid viscosity;
− density of solid particles;
− dimension of solid particles;
− morphology of solid particles;
− dimensions and geometry of the metallic component.

A simplified method for the evaluation of ER for monophase systems was proposed by Salama et al.
(Ref. /14/, Ref. /15/) and is summarised into the following expressions:

W × U2 × d
Monophase gas or liquid ER = 0.182
D 2 × ρm

Gas systems ER = 8 × W × U2/D2

where
− W = solid particles mass rate or sand production [kg/day]
− U = average flow velocity [m/s]
− d = solid particles size [microns]
− D = pipe diameter [mm]
3
− ρm = fluid density [kg/m ]

For water-oil systems the erosive wastage rate can be quantified for common oilfield geometries with
the procedure developed at the E/CRC of the Tulsa University (Ref. /19/, Ref. /20/).

For multiphase systems the solid particles are usually entrained in the liquid phase and the erosion
rate is strongly dependent on the flow regime , and more sophisticated models are used based on
flow pattern prediction as the DNV RP O501 model or the E/CRC model developed at the University
of Tulsa (Ref. /19/, Ref. /20/) or other proprietary software developed for this purpose.

The models for the prediction of the sand erosion rate use a base equation of the following type:
a
k ⋅ W ⋅ v p ⋅ F( ϕ )
ER =
A ⋅ ρt
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 46 of 51

where:
− K = factor expressing the effect of the material and the solid particles shape and type;
− Mi = solid particles mass flow rate impacting the erosion target area;
− vp = particle impact velocity;
− φ = particle impact angle;
− a = velocity exponent;
− A = erosion target area;
− ρt = target material density.

The main concern in predicting the erosion rate is in predicting the proper values of the solid particles
mass flow rate impacting a specific location of the inner pipe wall and of the impact angle and impact
velocity of the impacting particles. The term “direct impingement” applies when the mean direction of
particle motion intersects the pipe walls (such as in elbows). The term “random impingement” applies
when the direction of particle motion is parallel to the pipe walls (such as in straight piping). Erosion
rate due to direct impingement is usually an order of magnitude higher than that due to random
impingement.

5.8 Microbial Induced Corrosion (MIC)

Microbiological induced corrosion, MIC, in oil and gas industry is involved where superficial water is
used for operation, and in particular:
− water units as water injection, water disposal or firefighting;
− brackish water used for process purposes;
− brackish water used for temporary operations;
− water used during workover operation and muds for well perforation.

Oxygen free formation water is generally not promoting MIC, with the exception of tanks and barrels
(e.g. separators or decanters) not protected by anodes where MIC may occur at the vessel bottom.

The groups of bacteria met in oil and gas industry are:


− SRB - sulfate reducing bacteria: they are the most commonly met in anaerobic conditions, the
corrosion mechanism is cathodic depolarization;
− SOB - sulfate oxidizing bacteria: it is a group of bacteria growing in aerobic conditions and
leading to the production of sulfuric acid;
− IRB - iron related bacteria: it is a group of bacteria growing in aerobic conditions and leading to
the oxidation of iron ion from a soluble to insoluble, which in combination with chlorides give
place to corrosive ferric chloride;
− APB - acid producing bacteria are responsible for the formation of organic acids and sulfuric acid;
− ASB - aerobic slime bacteria may lead to the formation of "slime" that is a product with a
polymeric basis.

All aerobic bacteria consume the oxygen locally present and create the conditions for anaerobic
bacteria growth.

In the oil and gas industry, SRB is, the most common. Conditions for MIC by SRB include:
− presence of water;
− presence of bacterial activity (SRB);
− presence of sulfates (SO4-- above 10 ppm);
− anaerobic conditions (even only local);
− temperature below 80 °C;
− salinity lower than 140 g/l.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 47 of 51

The mechanism of SRB corrosion according to the cathodic depolarization theory involves the
following reactions (Ref. /21/, Ref. /22/, Ref. /23/):
Anodic reaction 4Fe → 4Fe2+ + 8e-
Water dissociation 8H2O → 8H+ + 8 OH-
Cathodic reaction 8H+ + 8e- → 8H(ads)
Depolarization mechanism SO4 + 8H → MIC → S2- + 4H2O
2+

Corrosion product Fe2+ + S2- → FeS


Corrosion product 3Fe + 6OH- → 3Fe(OH)2
2+

Global reaction 4Fe + SO42+ + 4H2O → FeS + 3Fe(OH)2 + 2OH-

Conditions for MIC initiation are:


− Presence of SRB.
− Presence of liquid water.
− Presence of sulphates ions for SRB growth: a minimum concentration of 10 ppm of sulphates is
assumed. SRB growth is also supported by thiosulphate ions formed by H2S and O2 reaction.
− SRB growth occurs in a temperature range from 0°C to a maximum of 70°C.
− SRB growth occurs in a pH range from 4 to 10.
− Anaerobic conditions: it is met mainly in deaerated fluids, or under deposits or inside interstices
in aerated fluids.
− Inside pipelines MIC preferentially occurs in stagnant sections or when flow velocity is low.

To avoid MIC, water is usually treated with biocides (chlorine is typically used for sea water).

No algorithms are available to quantify MIC corrosion rate. If MIC conditions are verified, the
following criteria may be applied to estimate corrosion rate vMIC:
− Fluid treated with biocides or chlorine and average fluid velocity ≥ 1: MIC could occur if the
treatment is not adequate or under deposits. In this case vMIC = 0.1 mm/y.
− Fluid treated with biocides or chlorine and average fluid velocity < 1: MIC is possible under
deposit or inside interstices, biocides are not able to reach the bacteria. vMIC = 0.5 mm/y.
− Fluid not treated with biocides or chlorine and presence of bacteria: vMIC = 1 mm/y.

5.9 Galvanic corrosion

Galvanic or bimetallic corrosion occurs when metallic materials displaying a different electrochemical
potential in a given electrolytic environment are coupled in electrical contact. On the less noble metal
- lower potential - accelerated anodic corrosion take place, whereas at the more noble metal the
cathodic reaction takes place.

Galvanic coupling gives place to an increase in corrosion rate or to localised corrosion only if the
environment has sufficient conductivity.

Galvanic contact effects can also occur between metal and scales, fragments or corrosion products
with electronic conductivity. It is the case, for instance, of magnetite scales or of iron sulfide corrosion
products, both showing a cathodic behaviour with respect to carbon or low alloy steel.

Damaging effects can be observed also on the more noble metal, if the hydrogen produced by the
cathodic reaction can cause hydrogen embrittlement to the more noble metal.

The following table summarises the different situations.


Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 48 of 51

Environment (T) Galvanic coupling effects


carbon steel - anode CRA - cathode
Ecorr less noble Ecorr more noble
low temperature − general corrosion − hydrogen embrittlement
(< 60 °C) − pitting and crevice
medium-high temperature − general corrosion − cathodic protection
(> 60 °C) − pitting and crevice
− stress corrosion

5.10 Elemental sulfur corrosion

5.10.1 Elemental sulfur from the reservoir

In hydrocarbons characterised by a very high presence of H2S, elemental sulfur may be produced by
the decomposition of polysulfides (H2Sx where x is equal or greater than 2) according to the following
reaction:
H2Sx → H2S + S(x-1)
Elemental sulfur is not normally detected in systems with low H2S concentration (lower than 10÷15%
mol).

Also oxygen contamination of hydrocarbon systems, naturally oxygen free, determines the risk of
elemental sulfur formation in H2S containing fluids. A slow reaction between oxygen and hydrogen
sulfide occurs at room temperature to produce elemental sulfur.
H2S + ½ O2 → H2O + S↓
At high temperature (above 150 °C) the reaction is kinetically favoured and elemental sulfur
precipitation is expected. Elemental sulfur can be very corrosive to most of the materials used in the
oil and gas industry. For this reason oxygen contamination shall be avoided.

In wet systems elemental sulfur can drastically increase corrosion in a cathodic reaction either
directly or through polysulfides.

Both carbon and low alloy steels and corrosion resistant alloys subjected to elemental sulfur
corrosion.

Carbon and low alloy steels.

The corrosion rate of carbon steel in presence of elemental sulfur is greatly influenced by the pH and,
above all, by chlorides presence. The deposits can also promote under-deposit corrosion and reduce
inhibition efficiency.

Sulfur is a strong oxidant: at temperature within 120 °C, elemental sulfur reacts with carbon or low
alloys steels, or with their oxides, to form sulfides.

At normal temperature, in anaerobic conditions and in presence of chlorides (above 1 g/l), general
corrosion rate becomes very severe, as high as 10 mm/y and higher, and increase with chlorides
content with a maximum around 35 g/l of chlorides. In such conditions the proposed mechanism is
cathodically controlled, involving autocatalytic formation of mackinawite and leading to catastrophic
corrosion rates. The mechanism by which iron sulfides as mackinawite catalyse cathodic processes
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 49 of 51

in sulfur containing systems is attributed to their good electronic conductivity, low overpotential for
hydrogen evolution, noble electrode potential and defect structure (Ref. /8/).

Corrosion rate increases with temperature. At temperatures above 120 °C corrosion rate increases
due to the formation of sulfuric acid (4S + 4H2O → 3H2S + H2SO4) and polysulfanes (H2Sx).

Corrosion resistant alloys.

CRAs in presence of elemental sulfur are susceptible to localised attacks and stress corrosion
cracking. Low nickel austenitic stainless steels (AISI 304) and duplex stainless steels are particularly
sensitive to elemental sulfur and corrosion rate is high as for carbon and low alloy steels.

Corrosion resistance to elemental sulfur increases in high alloyed material: nickel, chromium and
molybdenum have a beneficial effect; high nickel austenitic stainless steels, nickel alloys and cobalt
alloys can be susceptible to localised attack, as pitting, crevice or stress corrosion cracking, in
particular above 150 °C and when chloride ions are present.

Nickel-chromium-molybdenum alloys with Ni>50%, Mo>12% and Cr>15% are usually resistant to
elemental sulfur corrosion.

In presence of elemental sulfur, CRA materials suitable for this environment shall be selected in
conformity with requirements provided by ISO 15156-3.

5.11 Amine corrosion

Amine can produces a localized form of corrosion, which occurs principally on carbon steel in some
gas treating processes; however also stainless steels can suffer from this form of corrosion.

Generally, corrosion in MDEA and DGA solutions is less severe than in MEA and DEA when
contaminants are well controlled.

5.11.1 Carbon steels

In low pressure systems, corrosion of carbon steel can be most severe in units that primarily remove
carbon dioxide, that is, where the hydrogen sulfide content of the acid gas is less than 5% by volume.
Corrosion of carbon steel components has been least severe in units that remove only hydrogen
sulfide, and in units that handle mixtures of carbon dioxide and hydrogen sulfide.

In high pressure units with high hydrogen sulfide partial pressure, corrosion of carbon steel can be
severe.

Attack is most pronounced at locations where acid gases are desorbed (flashed) from rich amine
solution and where temperature and flow turbulence are highest.

Carbon steel corrosion in amine treating processes is a function of a number of interrelated factors.
API Publication 581 provides tables with indications of amine corrosion rate for combinations of the
following parameters.

Concentration of amine solution. The commonly used amine concentration values are: MEA~20%,
DEA~30%, MDEA~40÷50%. For higher amine concentrations the tabulated corrosion rate should be
multiplied by an appropriate factor ranging between 1 and 2.

Acid gas loading. It is defined as moles of acid gas per mole of active amine. ”Rich” solution is an
amine of higher acid gas loading than “lean” solution, which is typically lower than 0.1 mol/mol.
Corrosion in amines with high lean acid gas loadings is not an uncommon problem because lean
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 50 of 51

solution temperatures are often greater than rich solution temperatures. Both H2S and CO2 must be
measured to determine acid gas loading.

Heat Stable Amine Salt concentration (HSAS). These amine degradation products reduce the
amount of active amine available to absorb acid gas, resulting in higher acid gas loading. In addition,
some amine degradation products themselves are corrosive. In MEA and DEA systems, heat stable
amine salts above 0.5 wt% can increase corrosion although a common operating limit is 2 wt%.
Corrosion can be a serious issue, even at low acid gas loadings, if HSAS is above 2 wt%. MDEA
also forms heat stable amine salts but the primary influence on corrosion in these units is organic
acids contaminants. Corrosion in MEA units can be more severe than in those that use DEA,
because MEA is more prone to degradation. However, amine solutions such as DEA that are
normally not purified by reclaiming can also become quite corrosive.

Temperature. Higher temperature as per most of the corrosion mechanisms, increases the corrosion
rate.

Fluid velocity. Amine corrosion is fairly uniform if not enhanced by flow effects: high flow velocities
and turbulence could give place to acid gas evolution, resulting in localized corrosion. In addition,
high flow rates and turbulence can remove protective iron sulfide films increasing the corrosion rate.
For carbon steel, common velocity limits are about 1.5 m/s for rich amine and about 6 m/s for lean
amine.

5.11.2 Austenitic stainless steels

Austenitic stainless steels are selected for the most critical components in CO2 and H2S removal
units. These materials are less susceptible than carbon steel to amine corrosion and are commonly
used where corrosion rate of carbon steel is too severe. Typical high corrosion rate areas where
austenitic stainless steels are used include hot/rich solutions contact with high acid gas loading,
areas of high turbulence or fluid velocity, impingement, vapour flashing or two-phase flow and most
heat transfer surfaces operating above 110°C approximately.

Common applications for austenitic stainless steels are scrubbers, reboilers, reclaimers and hot
reach/lean amine exchanger tubes.

According to API Publication 581, for this class of materials amine corrosion rate only depends on
the acid gas loading. The highest corrosion rate for all amine types and concentrations is 0.15 mm/y
for acid gas loading above 0.7 mol/mol.
Eni S.p.A. 02555.VAR.COR.PRG
Rev. 3 May 2006
Exploration & Production Division
Pag. 51 of 51

6. BIBLIOGRAPHY

Ref. /1/ J.M.Mendhane, G.Gregory, K.Aziz “A Flow Patter Map for Gas Liquid Flow in Horizontal
Pipes: Predictive Model” Int.J.Multiphase Flow,
Ref. /2/ 1974P.Griffith, “Multiphase Flow in Pipes”, Journal of Petroleum Technology, 1984
Ref. /3/ Y.Taitel, A.E.Duckler “A Model for Predicting Flow Regime Transitions in Horizontal and
Near Horizontal Gas Liquid Flow”, AlChE, 1976
Ref. /4/ B.Kermani, J.Martin, K.Esaklul, “Materials Design Strategy: Effects of H2S/CO2 Corrosion
on Materials Selection” NACE Corrosion 2006, Paper.
Ref. /5/ M.Bonis, M.Girgis, K.Goerz, “Weight loss corrosion with H2S: Using past operations for
designing future facilities” NACE Corrosion 2006, Paper 06122.
Ref. /6/ B.Brown, K.Lee, S.Nesic “Corrosion in multiphase flow containing small amounts of H2S”
NACE Corrosion 2003, Paper 03341.
Ref. /7/ F.M.Pots et. al. “Improvements on deWaar-Milliams corrosion prediction and applications
to corrosion management” NACE Corrosion 2002, Paper 02235.
Ref. /8/ L.Smith, B.Craig “H2S and Sulfur Management in E&P Operation: Materials for Extreme
Sour Service”, Report prepared for ENI E&P, August 2004.
Ref. /9/ C.deWaard, U.Lotz, D.E.Milliams “Predictive model for CO2 corrosion engineering in wet
natural gas pipelines” NACE Corrosion 1991, Paper 91976.
Ref. /10/ C.deWaard, U.Lotz, D.E.Milliams “Prediction of CO2 corrosion of carbon steel” NACE
Corrosion 1993, Paper 93069.
Ref. /11/ C.deWaard, U.Lotz, A.Dugstad “Influence of liquid flow velocity on CO2 corrosion: a semi-
empirical model” NACE Corrosion 1995, Paper 95128.
Ref. /12/ M Wicks, J P Fraser, "Entrainment of watel by flowing oil", Materials Perfbmiance, Vol 14,
p 9, May 1975
Ref. /13/ A.Dugstad, L.Lunde, K.Videm “Parametric study of CO2 corrosion of carbon steel” NACE
Corrosion 1994, Paper 94014.
Ref. /14/ M.M.Salama, E.S.Ventakesh “Evaluation of API RP14E Erosional Velocity Limitations”
OTC 4485, 1983.
Ref. /15/ M.M. Salama “An Alternative to API RP14E Erosional Velocity Limits for Sand Laden
Fluids” OTC 8898, 1998.
Ref. /16/ U.Lotz “Velocity Effects in Flow Induced Corrosion”, NACE Corrosion 1990, Paper 90027.
Ref. /17/ E.Heitz “Chemo-Mechanical Effects of Flow on Corrosion” Corrosion, Vol.47, No.2, 1991.
Ref. /18/ J.S.Smart III “A Review of Erosion Corrosion in Oil and Gas Production” NACE Corrosion
1990, Paper 90010.
Ref. /19/ J.R.Shadley, S.A.Shirazi, E.Dayalan, E.F.Rybicki “Velocity Guidelines for Preventing
Pitting of Carbon Steel Piping when the Flowing Medium Contains CO2 and Sand”, NACE
Corrosion 1996, Paper 96015.
Ref. /20/ J.R.Shadley, S.A.Shirazi, E.Dayalan, E.F.Rybicki “Prediction of Erosion-Corrosion
Penetration Rate in a CO2 Environment with Sand”, NACE Corrosion 1998, Paper 98059.
Ref. /21/ NACE - TPC Publication 3, “Microbiologically Influenced Corrosion and Biofouling in
Oilfield Equipment”, 1990.
Ref. /22/ MTI Publication 13, “Microbiologically Influenced: a State-of-the-Art Review”, Second
Edition, Materials Technology Institute of the Chemical Process Industries, 1989.
Ref. /23/ S. Dexter, “Biologically Induced Corrosion”, NACE - 8, 1986.

You might also like