You are on page 1of 52

Manual for HYDROCOR 2001 spreadsheet for the prediction of corrosion in multi-phase

pipelines transporting wet hydrocarbons

by

B.F.M. Pots
1. Copyright....................................................................................................................4
2. Summary....................................................................................................................5
3. Introduction................................................................................................................6
3.1. What is available in this manual............................................................................................................................ 6
3.2. Who to contact...................................................................................................................................................... 6
3.3. Other reading material.......................................................................................................................................... 7
4. User's manual for standard application......................................................................8
4.1. Installation............................................................................................................................................................. 8
4.2. Running HYDROCOR........................................................................................................................................... 8
4.3. Description of input fields...................................................................................................................................... 9
4.3.1. Corrodents.................................................................................................................................................... 10
4.3.2. Operating conditions..................................................................................................................................... 10
4.3.3. Water analysis.............................................................................................................................................. 12
4.3.4. Corrosion control.......................................................................................................................................... 13
4.3.5. Pipeline......................................................................................................................................................... 13
4.4. Description of output fields.................................................................................................................................. 14
4.4.1. Maximum corrosion rate............................................................................................................................... 14
4.4.2. Maximum TOL corrosion rate....................................................................................................................... 14
4.4.3. Location of maximum corrosion.................................................................................................................... 14
4.4.4. Scale inhibitor failure.................................................................................................................................... 14
4.4.5. Flow pattern.................................................................................................................................................. 14
4.4.6. Liquid hold-up............................................................................................................................................... 15
4.4.7. Mixture velocity............................................................................................................................................. 15
4.4.8. Film velocity.................................................................................................................................................. 15
4.4.9. Relative film length....................................................................................................................................... 15
4.4.10. Watercut..................................................................................................................................................... 15
4.4.11. pH............................................................................................................................................................... 15
4.5. Graph curve selection......................................................................................................................................... 15
4.6. Storing and retrieval of cases.............................................................................................................................. 17
5. Special options for expert users...............................................................................19
5.1. Elevation profile option........................................................................................................................................ 19
5.2. Expert mode entries............................................................................................................................................ 20
5.2.1. Corrosion control.......................................................................................................................................... 20
5.2.2. Modeling....................................................................................................................................................... 20
5.2.3. Pipeline......................................................................................................................................................... 21
5.2.4. Water analysis.............................................................................................................................................. 21
5.2.5. Operations.................................................................................................................................................... 22
5.2.6. Fluid properties............................................................................................................................................. 22
5.2.7. Computational.............................................................................................................................................. 23
6. HYDROCOR limitations.............................................................................................24
6.1. Application limits................................................................................................................................................. 24
6.2. Accuracy of corrosion predictions....................................................................................................................... 24
7. Corrosion rate modeling in HYDROCOR....................................................................26
7.1. CO2 corrosion..................................................................................................................................................... 26
7.1.1. LCR+99 model............................................................................................................................................. 27
7.1.2. Comparison with lab data............................................................................................................................. 28
7.1.3. Top-Of-Line (TOL) corrosion......................................................................................................................... 31
7.1.4. Previous models........................................................................................................................................... 31
7.1.5. Corrosion product scaling............................................................................................................................. 36
7.1.6. Oil wetting..................................................................................................................................................... 38
7.1.7. Alcohol injection............................................................................................................................................ 38
7.1.8. Effect of total system pressure..................................................................................................................... 39
7.1.9. Corrosion inhibitor injection.......................................................................................................................... 39
7.2. Organic acid corrosion........................................................................................................................................ 39
7.3. H2S corrosion..................................................................................................................................................... 40
7.4. Bacterial corrosion.............................................................................................................................................. 41
7.5. Oxygen corrosion................................................................................................................................................ 43
8. Other essential models in HYDROCOR.....................................................................44
8.1. Water chemistry.................................................................................................................................................. 44
8.2. Alcohol/water distribution.................................................................................................................................... 44
8.3. Gas/liquid two-phase flow................................................................................................................................... 44
8.4. Pressure and temperature profiles...................................................................................................................... 44
8.5. Heat transfer model............................................................................................................................................ 45
9. MULTCASE examples................................................................................................47
10. Changes between versions....................................................................................48
10.1. WETGAS7.2/FLOWLINE7 & HYDROCOR1.04................................................................................................ 48
10.2. HYDROCOR1.04 & HYDROCOR2.01.............................................................................................................. 48
10.3. HYDROCOR2.01 & HYDROCOR3.01.............................................................................................................. 48
11. References..............................................................................................................49
1. Copyright

'Copyright is vested in Shell International Oil Products B.V., The Hague, The Netherlands. All rights reserved. Neither
the whole nor any part of this document may be reproduced, stored in any retrieval system or transmitted in any form
or by any means (electronic, reprographic, recording or otherwise) without the prior written consent of the copyright
owner.

This CONFIDENTIAL document is made available subject to the conditions that the recipient will neither use nor
disclose the contents except as agreed in writing with the copyright owner.

Although SHELL companies have their own separate identities the expressions SHELL and GROUP are used for
convenience to refer to companies of the Royal Dutch/Shell Group in general, or to one or more such companies as
the context may require.'

'Neither N.V. Koninklijke Nederlandsche Petroleum Maatschappij, nor The "Shell" Transport and Trading Company,
p.l.c. nor any company of the Royal Dutch/Shell Group will accept any liability for loss or damage originating from the
use of the information contained herein.'

SHELL INTERNATIONAL OIL PRODUCTS B.V.

SHELL GLOBAL SOLUTIONS, AMSTERDAM


Shell Global Solutions is a trading style used by a network of technology companies of the Royal Dutch/Shell Group
2. Summary

This manual explains how to install and use the Excel spreadsheet HYDROCOR 2001.

The spreadsheet comprises the latest upgrade of the prediction model for the rate of corrosion inside carbon steel
pipelines used for the transportation of gas and oil or condensate, containing corrosive, free water.

HYDROCOR contains corrosion rate prediction models for the main corrosion mechanisms of carbon steel in
Exploration and Production service. The models succeed earlier work by De Waard, Milliams, and Lotz.

Models are contained for the following mechanisms: CO2 corrosion, CO2/H2S corrosion, H2S corrosion, organic acid
corrosion, oxygen corrosion, and microbiologically-induced corrosion.

Application limits are indicated.

HYDROCOR 2001 or HYDROCOR version 3 supersedes the earlier spreadsheets WETGAS, FLOLINE, and the
HYDROCOR versions 1 and 2.

Houston, January 2002


3. Introduction

This is the manual for the HYDROCOR 2001 spreadsheet for the prediction of corrosion rates in multi-phase pipelines
transporting wet hydrocarbons.

HYDROCOR 2001 or HYDROCOR version 3 and subsequent v3.xx updates, is a PC-based engineering spreadsheet
for quantifying the corrosivity of the operating conditions associated with the transportation of water-wet hydrocarbons
in a carbon steel pipeline. With HYDROCOR the operational conditions can be identified under which a corrosive, wet
hydrocarbon production stream can be transported in a pipeline of carbon steel.

HYDROCOR, version 3 replaces all corrosion prediction software previously produced by Shell Global
Solutions, i.e. WETGAS, FLOLINE and HYDROCOR versions 1 and 2.

HYDROCOR 2001 is primarily intended to predict the corrosion rates profiles along pipelines. The program may also
be manipulated to afford corrosion rates for other EP equipment of similar geometry to pipelines e.g. tubing where
corrosion rate profiles are also required. With some imagination, the program can be used for other EP equipment,
including piping, vessels, heat exchangers, etc.

The program checks for the presence of free water considering the presence of water vapor in the gas phase,
dissolved water in the hydrocarbon liquid phase, and any free water entering either from production (formation,
condensation) or injection (alcohol). Based on whether free water is contacting the pipeline, corrosion rates are
calculated along the pipeline profile. This calculation requires a number of additional models, which are also
incorporated in the program.

Predictions from HYDROCOR provide the basis for assessment of the corrosivity of the service conditions. In addition
to such an assessment, consideration must be given to subsequent supplementary laboratory testing and field data.
Laboratory testing will be required to determine the corrosion rate under inhibited conditions and conditions outside
the application limits of HYDROCOR.

The remit of the HYDROCOR manual is to cover the use of the program to generate corrosion rates from available
input. The manual is not the document to cover the consequences of these corrosion predictions. These
consequences of e.g. corrosion inhibition (testing, selection, application), corrosion monitoring (FSM, ER probes),
inspection (intelligent pigging), risk-based integrity assessments (including RBI) and related operational requirements
(cleaning pigs) would need to be covered by an overall Corrosion Management document that in the broadest sense
covers all the prospect/project stages from feasibility through conceptual to detailed design to operations.

It remains the responsibility of the user to ensure that application of HYDROCOR is justified and that output results
are properly interpreted. Delegating HYDROCOR calculations to an inexperienced engineer or to outside contractors
without proper training is not recommended.

3.1. What is available in this manual

In this manual you can find:

What is new in the latest version (see changes between HYDROCOR 1999 and HYDROCOR 2001).
How to install HYDROCOR on your PC.
How to use HYDROCOR; an extensive description of all input and out paramaters for the Standard application
mode is provided.
How to store and retrieve scenario cases from a database spreadsheet.
How to use the special input fields for the Expert application mode (experienced users only).
How to simulate a full distance-elevation profile; the Standard mode deals with a horizontal pipeline only.
Information about the HYDROCOR limitations (deficiencies, uncertainties, arbitrary positions).
A description of the modelling behind HYDROCOR.
Example cases.

3.2. Who to contact

In case questions arise or potential bugs are found, please contact one of our specialists:
Amsterdam: Maarten Simon Thomas
Rijswijk: Ian Rippon
Houston: Bert Pots

3.3. Other reading material

An attempt has been made to provide all essential background information in this manual.
Other suitable reading material is:

Publication on CO2 corrosion design considerations in the European Federation of Corrosion report No. 23.
4. User's manual for standard application

4.1. Installation

HYDROCOR is an (Microsoft ) EXCEL spreadsheet and runs under Windows 95/98 and Windows NT.

The EXCEL version must be 7 or higher. HYDROCOR will not run under Windows 3.1.

HYDROCOR comes as a ZIP file (HYDROCOR2001.ZIP) and needs to be "unzipped" first. Four files are to be
extracted:

HYDROCOR 2001.XLS: spreadsheet to be opened;


MULTCASE.XLS: spreadsheet for storing/retrieving cases;
HYDROCOR2001 xyz.DLL: binary DLL file containing the prediction models.
HYDROCOR2001.HLP: helpfile containing this manual

The four files must be copied to one directory of (own) choice on the hard disk (e.g. C:\HYDROCOR 2001).
HYDROCOR can be run from any medium (diskette, CD, etc.), but copying ALL files to the hard disk into ONE
directory of own choice is recommended.

Minimum system requirements are a 486 IBM compatible PC with math co-processor running under Windows 95/98 or
Windows NT.

The program design philosophy of HYDROCOR is based on combining the advantages of user-friendly spreadsheets
and higher programming language libraries. Actual model calculations are not carried out in the spreadsheet but via a
Dynamic Link Library (DLL).

All source code is FORTRAN language. The DLL was produced from the source code using Microsoft's FORTRAN
compiler/linker POWERSTATION 4.0.

4.2. Running HYDROCOR

HYDROCOR must be started by first launching EXCEL. Thereafter, HYDROCOR must be opened from the directory
of choice (with "File", "Open" from the pull-down menu).

DO NOT START THE EXCEL SPREADSHEET BY CLICKING ON HYDROCOR 2001.XLS IN "EXPLORER" . FIRST
START EXCEL AND OPEN HYDROCOR 2001.XLS FROM THE DIRECTORY WHERE IT RESIDES. All other ways of
starting HYDROCOR are prone to fail as the associated DLL file may not be found.

Alternatively, the DLL can be copied to the C:\WINDOWS\SYSTEM directory. This will allow to start HYDROCOR from
the file manager as the DLL file will always be found.

The spreadsheet is to a great extent self-explanatory and can be used on its own. Extensive use was made of the
possibility to add notes to cells in the spreadsheet. These cells are marked with a red dot in the upper right-hand
corner.

The notes can be printed by selecting the menu: "File", "Page set-up", "Sheet", and ticking the box for the printing of
the comments.

Cases are best saved in the "database" file MULTCASE.XLS, see "Storing and retrieval of cases". This avoids the
need to save a copy of HYDROCOR 2001.XLS each time a new case is considered. The database file
MULTCASE.XLS can be saved under another name (to keep results for different projects in separate files). This
requires to retitle in sheet "Main" the MULTCASE.XLS name to the name of the renamed file and this file must reside
in the same folder as HYDROCOR 2001.XLS.

HYDROCOR is a workbook, which contains the following sheets:

MAIN: the main spreadsheet containing the main input and output screen. Changing one or more input
parameters will prompt recalculation of the spreadsheet after pressing the "RUN HYDROCOR" button.
ELEVATION: only to be used if an elevation profile is required;
EXPERT: only to be used by experienced users;
OUTPUT: shows detailed numerical output.

Calculations need to be initiated by clicking on the "RUN HYDROCOR" button in the sheet "Main".

There is no need to set Excel to manual recalculation before starting HYDROCOR.

4.3. Description of input fields

After opening HYDROCOR and leaving the opening screen, the spreadsheet "Main" will show up. This spreadsheet
contains the input and output fields for standard application.

Input/output screen of HYDROCOR

Cells in the spreadsheet are protected except for the blue input cells. Cells with a little red box in the right-upper
corner contain comments, which automatically will show when (slowly) moving over the cell by the mouse.

If a value outside the validity range is entered, no calculations will be performed and an error message results.

There are 4 blocks at the left-hand side for the input parameters:

CORRODENTS

OPERATING CONDITIONS

WATER ANALYSIS

CORROSION CONTROL.

PIPELINE

The entries in these blocks will be explained in detail in the subsequent sections.

In the upper-right part in sheet "Main", there are four check boxes:
1. H2S: ticking this box offers the option to deal with conditions with H 2S. For sour conditions (pCO2/pH2S<20), the
SOGACOR model of Shell Canada is used.
2. Imperial units: ticking this box changes the parameter units in sheet "Main" into imperial units.
3. Expert mode: ticking this box allows to change values of parameters in sheet "Expert". When un-ticked (Standard
mode), default values apply for the parameters in sheet "Expert", no matter what the (blue) values have been
entered. The special input fields for expert users are explained elsewhere in this manual.
4. Elevation profile: ticking allows to enter a pipeline elevation profile in sheet "Elevation". Also a variable heat
transfer coefficient along the pipeline can be taken.

4.3.1. Corrodents

The block "Corrodents" contains the species causing the corrosion.

4.3.1.1. CO2 content gas

Unit: mole%
Valid values: 0 - 100 mole%.

Enter CO2 mole percentage from the gas composition, if a gas cap is available.

For liquid-full lines, the CO2 gas content should be corrected for the difference between the pressure at the point of
last contact between gas and liquid and the actual inlet pressure.

For example: 1% CO2 in the gas in a separator at 10 bar(a) would give 0.1% CO 2 for an actual inlet pressure of 100
bar.

Alternatively, the entry cell "Dissolved CO2 in water" in sheet "Expert" can be used to enter the actual amount of
dissolved CO2 in the water phase (as may be available from an actual field measurement).

4.3.1.2. H2S content gas

Unit: mole%
Valid values: 0 - 100 mole%

Enter the H2S mole percentage from the gas composition. For the sweet and mixed corrosion regimes the
HYDROCOR CO2 corrosion model is used. For the sour corrosion regime the SOGACOR H 2S model is used.

The effect of H2S presence on the pH is included.

4.3.1.3. Free sulfur

Valid entries: Yes, No

Enter whether free elemental sulfur is present or not. Is only required for cases with H 2S, when in the sour regime.

4.3.2. Operating conditions

The block "Operating conditions" contains the following parameters:


4.3.2.1. Pressure inlet

Unit: bar or psi


Valid values: 1 200 bar(a)

Pipeline (absolute) inlet pressure.

4.3.2.2. Temperature inlet

Unit: C or F
Valid values: 0 150 C

Pipeline inlet temperature in C. For tubing calculations this entry relates to the bottom hole temperature.

4.3.2.3. Ambient temperature

Unit: C or F
Valid temperatures: 0 - 150 C

Temperature outside the pipeline. For an offshore pipeline this would be the sea water temperature. For buried land
lines this would be the soil temperature. For tubing calculations this entry relates to the well head temperature, which
requires the temperature model to be set to model No. 2 (linear profile) in sheet "Expert". For piping use the air
temperature.

4.3.2.4. Gas flow rate

Valid values: 0 500 mln Sm3/d

Gas flow rate in mln Sm3/d, where "S" refers to standard conditions, being 15 C and 1 atm. Flow rate needs to be
positive number. If the flow rate is too high for the pipeline diameter and length, the pressure will fall below 1 bara. In
this case a warning will be given and a lower gas flow rate needs to be entered (or a larger diameter).

4.3.2.5. Liquid HC type

Enter type of liquid hydrocarbons. This entry only affects the oil factor. For NGL's F oil=1 (no protection). Left to
judgement of corrosion engineer to choose between NGL's and oil (provides protection if water/oil forms dispersion).

An indicator for whether to take NGL or oil could be the API gravity, e.g.:

NGL's for API@60F>45

oil for API@60F<45

4.3.2.6. NGL/oil flow rate

Unit: m3/d or bbl/d


Valid values: 0 - 500,000 m3/d

Flow rate of liquid hydrocarbons. The current version of HYDROCOR does not include any phase equilibrium
calculation for the hydrocarbon components. If the liquid hydrocarbons flow rate would vary along the pipeline due to
evaporation/condensation this is not simulated . One can only enter a single value for which one could take the
average NGL/oil flow rate over the pipeline.
4.3.2.7. Water flow rate

Unit:m3/d or bbl/d
Valid values: 0 500,000 m3/d

Free water (aqueous phase) entering the pipeline as a discrete liquid phase, including: 1) free water carried through
knock-out vessel or separator, 2) free water from re-injected wet hydrocarbons, 3) alcohol injection, and 4) water
produced with the hydrocarbons directly from the well head. The water carried in the vapour phase should not be
included in this input cell as it is automatically included. Also water dissolved in the liquid hydrocarbons at the inlet
should not be included. The dissolved water property is included via sheet "Expert".

This flow rate includes any alcohol injection, if applied!

Water plus alcohols entering the pipeline as aqueous phase originating from, for example:
- alcohol injected ;
- carry over from knockout separator;
- FREE water in re-injected wet condensate;
- water produces with HC's directly from wellhead.

There are two other sources of water but they should not be inluded here because they are included in another way.
These sources are:
- water vapour in the gas phase;
- water dissolved in hydrocarbon liquids.

Glycol can be aplied to control corrosion. An example is Troll, where MEG is applied in the two 36" wet gas pipelines
to prevent both hydrate formation and high corrosion rates. The injection rate is 3 m3/mlnSm3 or120 m3/d at 40
mlnSm3/d per pipeline.

4.3.3. Water analysis

The block "Water analysis" contains the following parameters:

4.3.3.1. Water chemistry

Conditions refer to the pipeline inlet.

The following options are available:

1. No formation water: water does not contain water from the formation, only condensation water and water from
injected alcohol.
2+
2. Water at iron saturation: same as under item 1 except that Fe concentration is at FeCO3 saturation level. The
main effect is a (possible) increase of the pH.
3. Formation water: water from the formation is present, and may contain dissolved solids/acids prohibiting the
formation of a protective scale (Fscale=1)

4.3.3.2. Chlorides

Unit: ppm

Relevant for sour regime.

4.3.3.3. Bicarbonates

Unit: ppm
Bicarbonate level in total water comprising all sources of free water at inlet conditions. For evaluation of the pH control
option, the desired pH level can be adjusted via the bicarbonate level in this entry. Bicarbonates in formation water
should also be entered via this cell but need to be corrected for the fraction formation water. Valid values: 0 - 100,000
ppm (mg/l).

4.3.3.4. Organic acids

Unit: ppm
Valid values: 0-10,000 ppm

Organic acid concentration (as acetic acid) as measured in water. Note that organic acids are carried through the gas
phase.

4.3.4. Corrosion control

4.3.4.1. Annual wet operation

Percentage of time wet conditions exist. For a wet pipeline this will be normally 100%. For a nominally dry pipeline,
one needs to add the wet periods (dew point off-spec conditions, upset conditions leading to free water in the line, etc.
(including for example liquid glycol carry-over).

4.3.4.2. Inhibitor availaibility

Range: 0 100%.

Percentage of time that corrosion inhibitor is available. For conceptual design calculations an upper limit value of 95%
is deemed acceptable. Obviously, the actual design of the corrosion control system must afford this high availability to
be realised. Actual field experience shows availabilities may range between 50 and 99%

4.3.4.3. Alcohol type

Alcohols are applied to prevent hydrate formation in gas/condensate systems but will also reduce the water corrosivity,
thus offering a corrosion control option.

Type of alcohol injected (MEG, DEG, TEG, or methanol). Methanol option requires care w.r.t. division of pipeline into
sections. To avoid non-physical "oscillations", small section lengths may be required at the inlet (of the order of 10 m
or even less).

4.3.4.4. Alcohol concentation

Unit: w%

Alcohol concentration (weight) in total aqueous phase as entered in block "Water analysis". Typical composition of an injected
alcohol is 95% (w) pure alcohol and 5% (w) water. As other sources of water may be included, the alcohol concentration needs to
be corrected for the dilution by the other sources. Valid values: 0 - 99%. This entry allows controlling CO2 corrosion. MEG is
injected in the Troll pipelines (Norway) and brought corrosion down from 2 mm/y to 0.2 mm/y.

4.3.5. Pipeline

4.3.5.1. Pipe inside diameter

Unit: m or ft
Valid values: 0.001 - 2 m
If only outside diameter is available, subtract pipe wall thickness twice.

4.3.5.2. Pipeline length

Unit: km or mile
Valid values: 0.001 - 1000 km

In the Standard mode (i.e. "Expert mode" box unticked), a horizontal pipeline is assumed and the line is divided in 30
sections (separated by 31 nodes). Section lengths are automatically calculated by the program. The first section
length is 1m. Subsequent sections become longer by a factor of 10 1/5=1.6 until the pipeline length is reached.
Remaining nodes are used to halve the longest sections until all nodes are allocated. The short sections at the inlet
help to avoid numerical instabilities.

In case a point calculation is required (i.e. a 1-node calculation), the program should be run in the Expert mode and
the number of nodes in sheet "Expert" set to 1.

In case the box "Elevation profile" is ticked, this entry is overruled by the last node position (end of pipeline) provided
by the user.

4.4. Description of output fields

4.4.1. Maximum corrosion rate

Maximum CO2/H2S BOL corrosion rate found along the pipeline.

4.4.2. Maximum TOL corrosion rate

Maximum CO2 TOL corrosion rate found along the pipeline

4.4.3. Location of maximum corrosion

Location where the maximum corrosion rate is found (maximum of BOL or TOL). N.B. BOL and TOL maxima can
occur at different points but the location here applies to higher of BOL or TOL corrosion rate. All other output data in
sheet "Main" refers to this location.

4.4.4. Scale inhibitor failure

Failure due to exceeding critical (erosional) velocity of 20 m/s. Scale inhibitor failure is indicated by value of 1 (0 = no
failure).

4.4.5. Flow pattern

Two-phase flow pattern. Most relevant patterns are: stratified (liquid phase at bottom with gas phase on top of it),
annular dispersed flow (complete wall wetting by liquid film with part of liquid entrained as droplets in gas phase), and
intermittent or slug flow (liquid slugs followed by film zones).

Key for flow patterns:

1 = Stratified smooth
2 = Stratified wavy
3 = Stratified dispersed
4 = Annular
5 = Annular dispersed
6 = Intermittent/slug
7 = Churn
8 = Bubble
9 = Dispersed bubble
10 =Single phase flow

4.4.6. Liquid hold-up

Actual in-situ volume fraction of liquid. Normally, this fraction exceeds the inlet volume fraction of liquid due to slip of
the heavier liquid phase.

4.4.7. Mixture velocity

Sum of superficial liquid and gas velocities.

4.4.8. Film velocity

Actual velocity of liquid layer in stratified flow, liquid film in annular dispersed flow, or liquid film part of slug unit.

Film velocity can become negative for upward flow.

4.4.9. Relative film length

Length of film under slug flow. If length = 1 then flow is either stratified or annular dispersed. If length = 0 then flow is
(dispersed) bubble.

4.4.10. Watercut

Volume percentage of aqueous phase (including any alcohol) in total liquid phase (aqueous phase plus liquid
hydrocarbons.

4.4.11. pH

pH of water phase. Presence of any alcohol ignored in pH but included in corrosion rate.

4.5. Graph curve selection

The graph in sheet "Main", curves for two output parameters of choice can be shown.
By default the corrosion rate at the bottom of the line and the pH are selected. Other output parameters can be
selected via the drop-down forms at the bottom of the graph. Available output parameters are listed below:

List of available output parameters that can be selected for graph in HYDROCOR
CORROSION:
Corrosion rate bottom, mm/y
Corrosion rate top, mm/y
Corrosion rate film zone, mm/y
Corrosion rate slug zone, mm/y
Scale factor bottom
FeCO3 scaling tendency
Wall FeCO3 scaling tendency
Iron carbonate precipitation rate, mm/y
CO2 partial pressure, bar
Fugacity coefficient CO2
Oil wetting protection factor
Glycol factor bottom
Glycol factor top
Scale/inhibitor failure
Condensation factor top

FLUID FLOW:
Pipeline elevation, m
Pipeline section slope, degree
Superficial gas velocity, m/s
Superficial liquid velocity, m/s
Flow pattern Liquid hold-up (no unit)
Mixture velocity, m/s
Relative film length
Film velocity, m/s
Film depth, m
Film hydraulic diameter, m
Film friction factor
Slug hold-up (no unit)
Slug friction factor
Water cut
Water hold-up
Water film depth, m
Water-wetted perimeter, m
Liquid entrainment

PRESSURE & TEMPERATURE:


Pressure, bar
Temperature, C
Wall temperature, C

WATER CHEMISTRY:
pH
pH at wall
Saturation pH
FeCO3
2+
Fe , ppmw
2+
Fe concentration at wall, ppm
2+
Fe at saturation, ppmw
2+
Fe TOL corrosion, ppm
+ 3
Na , kmol/m

FLUID PROPERTIES:
3
Gas density, kg/m
2
Gas viscosity, Ns/m
3
Liquid density, kg/m
2
Liquid viscosity, Ns/m
2
Mixture viscosity, Ns/m
Surface tension, N/m
Heat capacity mixture, J/kgK
Heat conductivity mixture, W/mK

PHASE BEHAVIOUR:
Aqueous phase flow rate, kg/s
Water fraction aqueous phase
Vapour phase mass flow rate, kg/s
3
Water content gas phase, kg/mlnSm
3
Glycol content gas phase, kg/mlnSm
2
Condensation rate, g/m s
Water fraction condensing phase
2
Cold wall condensation rate, g/m s
Water fraction condensing droplets cold wall

4.6. Storing and retrieval of cases

The sheet "Load & Stores Cases" allows to load and store cases from/to the auxiliary database spreadsheet
MULTCASE.XLS. This avoids to have to save a copy of HYDROCOR each time a new case is considered.

Whilst the latest version of EXCEL allows the creation of user-friendly entry forms, HYDROCOR has been
programmed "leanly" in order to ensure that it will run on 99% of all machines, without the need to have special files
installed.

The new MULTCASE.XLS file is not compatible with the previous versions under HYDROCOR versions 1 and 2. With
cut-and-paste it is quite simple to convert from the old version to the new one.

A case (earlier stored) can be loaded from MULTCASE.XLS and run in HYDROCOR by:

selecting sheet "Load & Store Cases" in HYDROCOR;


opening the database MULTCASE.XLS by pressing the button "Open database"; this requires the working
directory to be the same as where MULTCASE.XLS resides;
selecting a case from the drop-down form "Select case";
loading and running the case by pressing the button "Load and run selected case";

A case can be stored from HYDROCOR in MULTCASE by:

selecting sheet "Load & Store Cases" in HYDROCOR;


pressing the button "Store case in database"

The case will be stored in the first empty column in MULTCASE


It is possible to run and store all cases in MULTCASE by pressing the button "Run and store all cases". The cases
will be stored starting from the first empty column in MULTCASE.
The modified MULTCASE can be saved by pressing the button "Save database". This should not be forgotten as
otherwise new cases are not saved and data will be lost.

Sheet "Load & Store Cases" for loading and storing cases from MULTCASE.XLS databases spreadsheet

It is possible to save a database under a different name than MULTCASE.XLS, which is convenient when one has to
deal with a number of projects. This requires to:

first change the name of the database in sheet "Main";


followed by pressing the button "Save database" in sheet "Load & Store Cases".

The alternative step 4 in sheet "Load & Store Cases allows to store a case over an existing case.
5. Special options for expert users

5.1. Elevation profile option

Pipeline elevation can have a dramatic effect on the two-phase flow pattern. Expert users have the option to enter an
actual pipeline elevation profile. Pipe sections can have different lengths and different angles of elevation The input
cells in sheet "Elevation" allow to enter an elevation profile in the form of a table containing distance along the
pipeline and the associated elevation. The maximum number of nodes is 101. The minimum number is 1. The first
row refers to the node at the pipeline inlet.

For a horizontal line it suffices to fill in the total pipeline length in the second row, first column. As there is no elevation,
the second column needs to contain a zero.

The total number of nodes used by HYDROCOR remains controlled by the corresponding cell in sheet "Expert",
unless the number in this entry is less than the number of nodes asked for in the elevation profile, in which case the
latter number applies.

A variable heat transfer coefficient can be entered, which is in effect if the heat transfer coefficient in sheet "Expert" is
set to zero. This allows the user to consider the effects of changes in pipeline coatings/concrete along the length of the
2 2
pipeline and riser. If the coefficient in column D is less than 0.01 W/m K, the default value of 20 W/m K is taken.

Elevation profile (SI units only!)


Node Distance Elevation Heat transfer
2
m m W/m K
1 0 0 20
0
2 5 -5 20
0 1000 2000 3000 4000 5000 6000
3 10 -10 20 -20
4 20 -20 20
Elevation, m

-40
5 30 -30 20
-60
6 40 -40 20
7 50 -50 20 -80

8 75 -75 20 -100
9 100 -100 20
-120
10 125 -100 20 Distance along pipeline, m
11 150 -100 20
12 200 -100 20 The red markers in above graph are the input pipeline nodes as provided by the user in
13 5000 0 20 colums B and C of this sheet. Blue markers are computational nodes, which are
14 0 0 20 (automatically) added by HYDROCOR. The total number of nodes is set by the value in cell
C39 in sheet "Expert" (maximum is 101 nodes).
15 0 0 20
16 0 0 20
When the elevation profile is active (i.e. corresponding box ticked in sheet "Main"), it
17 0 0 20 overrides the pipeline length entry in cell C7 of sheet "Main".
18 0 0 20
19 0 0 20 Elevation can be positive or negative.
20 0 0 20
21 0 0 20 First node can start with (x,y) values different from (0,0).
22 0 0 20
23 0 0 20 Notes regards choice of pipeline section length values:
24 0 0 20
25 0 0 20 1. Wiggly or jumpy curves in the "Main" graph indicate that too large section lengths are
taken in the elevational profile input.
26 0 0 20
2. The maximum BOL/TOL position values cannot be more accurate than the inputted
27 0 0 20
elevation distances.
28 0 0 20 3. Flow regime changes depend on the elevational profile. Pipeline section division should
29 0 0 20 reflect significant profile details.
30 0 0 20 4. An inlet/outlet riser should be represented by sufficient nodes (say at least 5 to 10
31 0 0 20 nodes).
32 0 0 20
33 0 0 20 An advised example input for a 100-m inlet riser is:
34 0 0 20
35 0 0 20 distance: 0, 5, 10, 20, 30, 50, 75, 100, 150, 200, 300
36 0 0 20 elevation: 0, -5, -10, -20, -30, -50, -75, -100, -100, -100, -100
37 0 0 20
To avoid wiggly output, inlet section lengths should be shortened until wiggles dissappear.
38 0 0 20
Application of methanol injection is a notorious example.
39 0 0 20
40 0 0 20
41 0 0 20
42 0 0 20
43 0 0 20
44 0 0 20
In-between rows not printed

96 0 0 20
97 0 0 20
98 0 0 20
99 0 0 20
100 0 0 20
101 0 0 20
Sheet "Elevation" in workbook HYDROCOR, simulating offshore 5-km pipeline with upstream riser of 100m
5.2. Expert mode entries

For advanced users there are a number of special input cells in sheet "Expert". Note that the values in these cells are
"overruled" by the default values when in the Standard mode. Only in the Expert mode do the (blue) values.

A number of entry parameters in the expert mode are related to other mechanisms than CO 2 or organic corrosion,
namely bacterial (MIC) and O2 corrosion.

5.2.1. Corrosion control

5.2.1.1. Inhibited corrosion rate

Default: 0.1 mm/y

Residual corrosion rate during periods that inhibitor is applied and effective. 1996 MCSM paper F7 suggests 0.2 mm/y as default value. This is
conservative. Therefore, the later CO2 corrosion position papers recommend 0.1 mm/y up to 120 C (0.3 mm/y up to 150 C) Lower values can be
entered based on relevant field, laboratory or project specific test data

5.2.1.2. Dewpoint pressure

Relevant entry for dry gas lines.

5.2.1.3. Dewpoint temperature

Relevant entry for dry gas lines.

5.2.1.4. Oxygen ingress

Unit: % of time

Relates to O2 corrosion.

5.2.1.5. Biocide routinely used

Valid entries: 0=yes, 1=yes

Relates to MIC.

5.2.2. Modeling

5.2.2.1. Corrosion model

Default: 5 (LCR+99)
Valid values: 1 (NACE91), 2 (IFE), 3 (LCR), 4 (LCR+), and 5 (LCR+99)

5.2.2.2. TOL corrosion model

Default: 4=model D (supersaturation model)

Other valid value is 2=Model B (De Waard/Milliams nomogram plus condensation factor based actual water condensation rate.

5.2.2.3. Temperature profile model

Default: 1 (Heat transfer model)

There are two temperature profile models:


1. Heat transfer model.
2. Linear temperature model.
In the latter case the ambient temperature entry in cell C10 in sheet "Main" is used for the pipeline outlet temperature. For tubing applications,
model 2 needs to be taken. So for tubing we adopt the flowing wellhead tubing temperature (FWHTP) at the top of the tubing and linearly "plotting"
the temperature profile from the reservoir temperature to the top of tubing.
5.2.3. Pipeline

5.2.3.1. Pipeline inclination

Pipeline inclination if check box "Elevation profile" in sheet "Main" profile is NOT ticked, otherwise this input is void. Positive numbers refer to
upflow. Valid values : -90 to +90

5.2.3.2. Pipe wall heat transfer coefficient


2
Unit: W/m K

2
Heat transfer coefficient of pipeline to outside environment. Value of 20 W/m K applies for a typical non-buried concrete-weight-coated seabed
pipeline with 24 mm steel wall thickness, 5 mm asphalt coating, and 50 mm concrete cover (see this table). When a temperature profile would be
2
available, the coefficient may be adjusted to fit this profile. Valid values: 0.01 10,000 W/m K.

5.2.3.3. Pipe wall roughness

Default: 0.07x10-3

Refers to hydraulic wall roughness relevant for frictional pressure drop. May be adjusted to get desired pipeline outlet pressure. Valid values: 0.001
- 1 mm.

5.2.3.4. Age of pipeline

Relates to MIC.

5.2.4. Water analysis

5.2.4.1. Dissolved CO2 in water

Default: 0 ppm

Amount of dissolved CO2 in water. This entry may be convenient for liquid-full pipelines with no free gas phase. This requires a measurement of the
amount of dissolved CO2 in the water to be available. It avoids the problem of having to interpret the mole% CO 2 in a gas phase upstream of the
pipeline. If the dissolved CO2 can only be measured in the oil phase, one needs to calculate the dissolved CO2 in the water phase from the
partitioning of the CO2 between the two phases. The partitioning can follow from a flash calculation by a process engineer.

5.2.4.2. Fe2+ concentration

At inlet conditions.

This option allows to simulate the effect of an enhanced ferrous ion concentration at the pipeline inlet due to an upstream corrosion process. Valid
values: 0 - 10,000 ppm.

5.2.4.3. In-situ oxygen concentration

Relates to O2 corrosion.

5.2.4.4. Sulphates

Unit: ppm

Relates to MIC.

5.2.4.5. Nitrogen (as N)

Unit: ppm

Relates to MIC.

5.2.4.6. Total dissolved solids

Unit: g/l

Relates to MIC.
5.2.5. Operations

5.2.5.1. Demulsifier injected

Valid values: 0(no), 1(yes)

If demulsifier is injected, it will be assumed that water will drop out.

5.2.5.2. Debris in pipeline

Valid values: 0(no), 1(yes)

Relates to MIC.

5.2.5.3. Routine pigging interval

Unit: wk

Relates to MIC.

5.2.5.4. Total operational downtime

Unit: wk

Relates to MIC.

5.2.6. Fluid properties

5.2.6.1. Gas molecular weight

Default value: 18 kg/kmol


Valid values: 16 - 100 kg/kmol

To be derived from gas composition or flash program.

5.2.6.2. Gas compressibility factor

Default value: 0.9


Valid values: 0.3 - 2

To be derived from flash program.

5.2.6.3. NGL/oil density


3
Unit: kg/m
3
Valid values: 500 - 1500 kg/m

NGL/oil density at prevailing line conditions. If nothing else is available on may consider to use

60/60=141.5 / (API-131.5) x 1000

where

3
60/60 is the density in kg/m .

5.2.6.4. NGL/oil viscosity

Default: 0.001Ns/m2
-3 2
Valid values: 0.01 10,000 cP (1cP=10 Ns/m )

NGL/oil viscosities cover a wide range of values. Therefore, the actual oil viscosity needs to be provided by the user. Reference temperature is
25 C. If nothing else is available, one may use the following API gravity "correlation" (ball park figures):

API viscosity, cP
----------------------------
>50 1
40 3
30 10
20 100
15 1000
5.2.6.5. Gas/liquid surface tension

Default: 0.05 N/m

0.001 - 0.1 N/m. Water surface tension is about 0.07 N/m (70 dyne/cm). Condensates can have surface tensions as low as 0.005 dyne/cm.

5.2.6.6. Water solubility NGL/oil C1 constant

Default: 7.9

Parameter C1 for solubility of water in NG/oil. Solubility is described by relation:

10
log (solubility in ppmw)=C1-C2/T

where T is the absolute temperature (Kelvin).

The water solubility parameters allow simulating the condensation of dissolved water from liquid hydrocarbons. This is
relevant for situations where, for example, condensate has been dried at a platform but becomes wet at lower (sea
bed) temperatures because of condensation of the residual (physically) dissolved water. A process engineer can
calculate the saturation water level of liquid hydrocarbons from thermodynamic flash (or pVT) models. This
information can then be used to adapt the C1 and C2 coefficients. This can be done by plotting the logarithm of the
water solubility in ppmw versus the reciprocal absolute temperature. The coefficient C 1 is the intercept and the C2
coefficient the slope of the linear trendline through the data points. The current values describe SSB (Sarawak)
condensate from the M fields.

5.2.6.7. Water solubility NGL/oil C2 constant

Default=1700

Parameter C2 for solubility of water in NGL/oil.

5.2.7. Computational

5.2.7.1. Number of pipe nodes

Default = 31

Number of computational nodes along pipeline = number of pipe sections plus one (nodes refer to the interfaces between the pipe sections). The
maximum number of nodes is 101, the minimum 1. If check box "Elevational profile" is ticked, this input is allied to sheet "Elevation", see text box in
this sheet. For N=1 calculations are of the type of the former program FLOLINE7.

5.2.7.2. Parameters shown for node

In the default mode, output parameters in sheet "Main" refer to the position where the maximum corrosion rate is found. If another value than zero is
taken, the corresponding node number is taken. Valid values: 0 101
6. HYDROCOR limitations

6.1. Application limits

Application limits may be software controlled, i.e., entry of data outside the application limits may give an error.

Application limits of HYDROCOR

Presence of H2S is included.

6.2. Accuracy of corrosion predictions

The accuracy of the HYDROCOR corrosion rate predictions depends on the application and the service conditions.
Find below an overview of the accuracy to be expected for a number of applications and conditions. More work is
needed for getting a more complete picture.

Accuracy of IFE and LCR+99 models for a number of applications and service conditions.
+25% means 25% overprediction, -25% means 25% underprediction
7. Corrosion rate modeling in HYDROCOR

The ability to confidently predict the internal corrosion of E&P carbon-steel facilities is essential for both front-end
design materials engineering and for managing lifetime integrity through optimum corrosion control.

HYDROCOR is more than a simple corrosion rate predictor. Additional models are included for:

water chemistry (pH calculation, iron supersaturation, FeCO 3 precipitation);


multi-phase flow (flow pattern, liquid hold-up, pressure gradient, phase flow velocities);
temperature gradient;
water/alcohol phase distribution;
water/alcohol condensation.

There are basically two types of CO2 corrosion modes:

corrosion of the bare steel without protection from a corrosion-product scale layer;
corrosion under scaling conditions with protection from a scale layer.

The program aims at evaluating whether corrosion is under non-scaling or scaling conditions.
A distinction is made between corrosion caused by the bulk of the water phase and corrosion under dewing conditions
in the top of the line under stratified flow conditions.

Calculation of the corrosion by the bulk of the water phase is based on evaluation of the bare-steel corrosion rate,
which rate is multiplied by a number of factors to account for the effects of total system pressure, corrosion product
scaling, oil wetting, alcohol injection, and inhibitor injection. For top-of-line (TOL) or dewing corrosion a separate
model is utilised.

The bare-steel corrosion rate is the corrosion rate that prevails when the corroding surface is not covered by a
corrosion product (layer).

Bare-steel corrosion rate models

Four models are available for the bare-steel corrosion rate:

(a) The De Waard/Lotz/Milliams model according to the 1991 NACE paper [2] based on the de Waard/Milliams
nomogram and including the pH factor in this paper. Although the model has been superseded by more recent
Shell models, the model is considered as one of the industry standards and users may want to use it as a
reference. Outside Shell this model may be referred to as "Shell 1991" (the "Shell 1993" is almost identical and
not separately considered).
(b) The de Waard et al. model based on the resistance concept and the IFE lab data from the KSC-II project [3-5]
the so-called 1995 IFE fit. IFE stands for Institutt for Energiteknikk (Norway). Outside Shell, this model may be
referred to as "Shell 1995".
(c) The Limiting Corrosion Rate Plus (LCR+) model, which is a mechanistic model based on fundamental relations
for species mass transport, chemical reaction equilibria and kinetics, and electro-chemistry [6].
(d) The 1999 version of the LCR+ model, which includes a number of improvements reducing conservatism in the
LCR+ model. A more detailed description of the LCR+99 model can be found elsewhere in this manual.

The main parameters for the bare-steel corrosion rate are: partial pressure of CO 2, temperature, pH, and flow. The
main flow parameters are the liquid flow velocity and the liquid hydraulic diameter.

The recommended bare-steel corrosion rate model is the LCR+ 1999 model.

Steel composition and microstructure are parameters in the 1995 IFE fit but may introduce unnecessary complexity.
Therefore, default values are provided in the program (normalised steel with 0.1 w% carbon and 0.05 w% chromium).
These parameters are not part of the LCR+ models.

7.1. CO2 corrosion

Different models are needed for Bottom-Of-Line (BOL) or bulk-water corrosion and Top-Of-Line (TOL) or dewing
corrosion. In previous models, a factor was introduced in the bare-steel BOL corrosion rate for the prediction of dewing
corrosion at stratified wavy flow conditions. This approach was abandoned as it was found that TOL corrosion bears
no simple, direct relation with the bare-steel BOL corrosion rate.

For TOL CO2 corrosion, the improved approach is recommended. This approach assumes that, at scaling conditions,
2+
the TOL corrosion rate is controlled by the Fe super-saturation level that is required in the condensation water to
maintain the iron carbonate scale. The main model parameter is the water condensation rate. Good agreement of the
model with both lab and field data was obtained. It should be noted that high-quality TOL corrosion lab tests need
durations of at least a number of months.

The starting point for BOL corrosion rate prediction is the bare-steel corrosion rate, which corrosion rate is multiplied
by factors to account for the effects of scale protection, oil wetting, and glycol/methanol presence. CO 2 fugacity rather
than partial pressure is used as a measure for the CO 2 activity.

7.1.1. LCR+99 model

The LCR+99 is an improved version of the LCR+ model. This improved version was made to eliminate some of the
conservatism (i.e. overprediction) in the 1995 version.

The following adaptations were made in the coefficients of the LCR+ model:

The temperature dependency was removed above 80 C for the:


electrochemical kinetics of the oxidation/reduction reactions (i.e., the rate is not further increasing above 80
C);
CO2 hydration rate.

The CO2 hydration rate was reduced (the rates are still within the scatter band seen in literature data).

0.7
The CO2 dependency was changed from linear to pCO2 in the
electrochemical reduction of carbonic acid;
CO2 hydration.

The proton dependency in the electrochemical reduction of protons was changed from linear to quadratic.

After introduction of above adaptations, an improvement was oberved in the agreement between model and
experimental data (see "Comparison with lab data")

Following the De Waard-Milliams-Lotz approach a more mechanistic approach for corrosion rate prediction was
developed starting in 1995. In this approach, species transport and reactions in the flow and reaction boundary layers
near the wall, as well as electrochemical reactions at the wall are considered using first principles. The Limiting
Corrosion Rate (LCR) model presented in Ref. forms the basis for describing the species transport and reactions. The
LCR model gives the limiting corrosion rate as set by the mass transport of corrosive species to the wall and the
chemical reactions producing the aggressive species. The extension presented in this paper involves the combination
of this LCR model with basic equations of electrochemical reaction kinetics following the mixed-potential theory:

k i ci e E / bi
j k Fe c Fe e E / bFe
i 1, 2 k c e E / bi
1 i i
j lim,i

The first term represents the anodic current related to iron dissolution and the second term the total cathodic current;
the temperature-dependent k's are the charge transfer reaction rate constants, the b's the Tafel slopes, c's the
concentrations (cFe=1), and jlim's the limiting currents from the LCR model. Protons (i=1) and carbonic acid (i=2) are
assumed to be cathodically active. At the corrosion potential E=Ecor, the anodic and cathodic currents are equal (j=0),
and equal to the corrosion current. The model is referred to as LCR+ model.

For the charge transfer kinetics, rule-of-thumb data for iron dissolution (b=40 mV/decade) and proton or carbonic acid
reduction (b=120 mV/decade) were taken. It was assumed that charge transfer for protons and carbonic acid are the
same. This leaves only one empirical parameter in the model, which is one reaction rate constant k plus its Arrhenius
coefficient. Equal anodic and cathodic k constants can be used, as the zero potential is arbitrary. The electrochemical
reaction rate constant was determined from glass-cell potentiodynamic curves.
7.1.2. Comparison with lab data

The new LCR+ CO2 corrosion model was compared with "high-quality lab" data generated at IFE (Institute for Energy
Technology, Norway) in the Kjeller Sweet Corrosion projects and at the authors' Amsterdam lab. Only data were
considered where scale formation was absent. High quality means in particular that 1) the solution chemistry is
controlled, particularly the iron level, and 2) the flow defined. On top of this quality requirement, it was ensured that
experimental data points were evenly distributed over the parameter space. Assuming that pCO 2, temperature, pH
and flow velocity are the main parameters, the 4-dimensional space made up by these parameters was divided in 72
windows or categories. For each category a maximum of 3 experimental data points was allowed.

The following process was followed to check the LCR+99 model:

Assuming pCO2, T, pH and flow velocity are the main parameters, the 4-dimensional space made up by these
parameters was divided in 72 categories, see Table 1.
For each category a maximum of 3 experimental data points was allowed.
Only high quality data points were used, see Table 2.
Only data referring to non-scaling conditions were considered (i.e. bare steel).

Above approach avoids comparison of lots of data only in limited parts of the parameter space (i.e. it is quite tough a
test).

Table 1. Division of parameter space


-----------------------------------------------------------------------------------------------
pCO2 <0.2 bar Medium >5 bar
Temperature <20 C Medium >60 C >100 C
pH Unbuffered Buffered
Flow velocity <0.2 m/s Medium >5 m/s
------------------------------------------------------------------------------------------------

Table 2. Experimental data used to validate models


------------------------------------------------------------------------------------------------
Data Comments
------------------------------------------------------------------------------------------------
IFE KSC-I tests
IFE KSC-II tests, Ref. 3 Both buffered and unbuffered
IFE KSC-III tests, Ref. 13 Buffered
IFE KSC-IV tests, Ref. 16 Buffered
IFE KSL-V tests, Ref. 14 Buffered
Gerretsen et al. 1993, Ref. 15
High temperature tests for B11 by IFE, Ref. 8
Original de Waard/Milliams data At iron saturation
------------------------------------------------------------------------------------------------

Above approach avoids comparison of lots of data only in limited parts of the parameter space (i.e. it is quite a tough
test).
Figures 3-1 to 3-4 compare models and experimental data. The new LCR+99 model performs best.
Figure 3-5 shows a comparison of the models with laboratory corrosion rates as a function of temperature, under the
most severe, high flow conditions available. Although data points may correspond to different CO 2 partial pressures,
0.7
corrosion rates were normalised to pCO2=1 bar (using CR~pCO2 ). The graph demonstrates:
Above 80 C, the worst-case corrosion rates are not further increasing.
The LCR+99 model no longer is so conservative above 80 C.
Actual corrosion rates can still be very high above 80 C (The data points at 150 C are probably under scaling
conditions but included as no zero no-scaling cases are available at this high a temperature).
1000

100
Predicted corrosion rate, mm/y

10

NACE91 model
0
0.1 1 10 100 1000
Measured corrosion rate, mm/y
Figure 3-1
Comparison of predicted CO2 corrosion rates of NACE91 model with measured laboratory data. Upper and lower lines
refer to +50% and - 33% deviation

1000

100
Predicted corrosion rate, mm/y

10

IFE fit
0
0.1 1 10 100 1000
Measured corrosion rate, mm/y
Figure 3-2
Comparison of predicted CO2 corrosion rates of IFE fit with measured laboratory data
1000

100
Predicted corrosion rate, mm/y

10

LCR+ model
0
0.1 1 10 100 1000
Measured corrosion rate, mm/y
Figure 3-3
Comparison of predicted CO2 corrosion rates of LCR+ model with measured laboratory data

1000

100
Predicted corrosion rate, mm/y

10

LCR+99 model
0
0.1 1 10 100 1000
Measured corrosion rate, mm/y
Figure 3-4
Comparison of predicted CO2 corrosion rates of LCR+99 model with measured laboratory data
Corrosion rate, mm/y 100

10
KSC-I 8.5-20 m/s
KSC-II 8.5 & 13 m/s
KSC-III 6.8 & 9.4 m/s
KSC-IV 6.8 m/s

KSC-V 6.8 m/s


B11 5 m/s
1 De Waard et al. 1991
model
De Waard 1995 model

LCR+ model
LCR+99 model

0
0 20 40 60 80 100 120 140 160

Temperature, C
Figure 3-5
Measured and predicted CO2 corrosion rates versus temperatures for high-flow conditions; data normalised to
pCO2=1 bar; calculated lines refer to unbuffered solution, V=10 m/s, D H=0.05 m (data points may refer to buffered
conditions)
Comparison of predicted CO2 corrosion rates of LCR+99 model with measured laboratory data.

7.1.3. Top-Of-Line (TOL) corrosion


.
Top-of-line (TOL) or 12 o'clock corrosion is considered for stratified flow conditions Prediction of TOL corrosion in
HYDROCOR version followed from multiplying the "de Waard/Milliams nomogram" corrosion rate with a condensation
factor, which takes account of the low water loading in dewing corrosion
G
Fcond
2.5
2
where G is the water condensation rate in g/m s. The nomogram is used because it was the basis for the development
of the above condensation factor.

The same fugacity coefficient is used as for bottom-of-line (BOL) corrosion.

The scale and oil factors are not applied in TOL corrosion. The condensation factor already reflects the effect of scale
protection and oil is assumed not to see the top of the line.
The alcohol factor is applied. Differences in alcohol concentration in the condensing droplets at the top and the water
in the bottom are considered (see water/alcohol phase distribution model in Section 2.4).

Corrosion inhibitor is assumed not to reach the top of the line under stratified flow conditions.
TOL corrosion model D from Ref. 12 was implemented and is considered to be the default.

7.1.4. Previous models

7.1.4.1. Overview

Over the years, various CO2 corrosion models have been developed by Shell:

"De Waard-Milliams" nomogram

Extensions of the nomogram by de Waard, Milliams, and Lotz (NACE 1991 and 1993 papers), outside Shell
often referred to as the Shell 1991 and 1993 models

De Waard et al. resistance model based on the IFE KSC-II project data, referred to as the IFE fit (or Shell 1995
model)

LCR model (NACE 1995 paper by Pots)

LCR+ model (1995)

LCR+99 model (1999)

7.1.4.2. De Waard-Milliams-Lotz based models

This model is only mentioned here for reference purposes. It should not be used any longer as it has been
superseded by LCR+99 model.

The original De Waard-Milliams nomogram equation was developed in the seventies from a series of stirred
autoclave experiments. The only two parameters were the CO 2 partial pressure and the temperature. As no clear
evidence for the effect of flow was found at the time, the "deWaard-Milliams" model assumed corrosion to be
controlled by the kinetics of the corrosion reactions. It ignored the aspect of mass transport of corrosive species.

Later, a correction factor was introduced to take account of the effect of the pH. The original nomogram assumed a
saturated iron carbonate level in the aqueous solution. Higher corrosion rates may be experienced for pH's lower than
the pH at saturation, however. This is relevant for situations where corrosion has no "opportunity" to build up a
sufficiently high iron level, such as at pipeline inlets with no upstream corrosion process or for wet gas systems with
high water condensation rates. Lower pH's can also result from the presence of other acids than carbonic acid.

Experiments at IFE, Norway, in the early nineties, demonstrated that flow velocity is an important parameter,
particularly for low pH's. At iron carbonate saturation conditions, the effect of flow is small, however, which explains
why no flow effect was observed in the De Waard-Milliams autoclave experiments.

7.1.4.3. IFE fit

Semi-empirical model based on IFE KSC-II project data, referred to as IFE fit. Shows good fit with these IFE data but
underprediction occurs for low flow velocities. For this reason the correlation should not be used below liquid flow
velocities of 0.5 m/s.

When it became clear from both in-house data and experiments at IFE (Institutt For Energiteknikk) in Norway, that
the flow rate can have a significant effect on CO2 corrosion and that corrosion rates can be much higher than
calculated from "de Waard-Milliams", a new semi-empirical model was developed based on the resistor concept. In
this resistance model, corrosion is controlled by having mass transfer and charge transfer resistances in series.
The mass transfer part contains the effect of fluid velocity. Allowing for five adjustable constants, the model was
fitted to data of IFE from their KSC-II project. The data refer to single-phase flow conditions.

7.1.4.4. LCR+ model

The LCR+ model is an extension of the LCR model. The extension refers to the incorporation of the electro-
chemical kinetics of the cathodic and anodic reactions.

The LCR+ model is based on the mixed-potential theory, which implies that at the corrosion potential (E=E cor), the
anodic and cathodic currents are equal (i.e. the net current density is zero):

The anodic current (i.e. the iron dissolution) is:

where ka is the temperature-dependent charge transfer reaction rate constant, b a the Tafel slope (40 mV/decade),
and E the potential.

The cathodic current for the electro-chemical reduction of species i is written as:
where ci is the concentration and jlim,i the limiting current as obtained from the LCR model (b c,i=120 mV/decade).
Protons (i=1) and carbonic acid (i=2) are assumed to be cathodically active species.

The k's for the cathodic reactions are assumed to be the same. This leaves only two empirical parameters in the
model, which are one rate constant k and its Arrhenius coefficient to describe the temperature dependence. Note
that the zero level of the potential is arbitrary, allowing equal anodic and cathodic k constants.

After solving the current equation (j(E cor)=0), the corrosion rate follows directly from the anodic current j a.

7.1.4.5. LCR model

Mechanistic model for the prediction of CO2 corrosion rates


INTRODUCTION

The Limiting Corrosion Rate (LCR) model is a mechanistic model assuming the rate-determining factors are the
hydration kinetics of CO2 and the mass transport of the corrosive species (protons and carbonic acid) to the steel
wall. The LCR model provides a theoretical upper limit for the corrosion rate.

The model needs no calibration with experimental corrosion data. A shortcoming of the LCR model is that it only
provides an upper limit for the corrosion rate. It therefore never functioned as a recommended model for corrosion
rate prediction within Shell (only its successors, the LCR+ and LCR+99 models did). For many applications the
corrosion rate will be too conservative, particularly for high-velocity situations when charge transfer kinetics instead
of mass transfer becomes the controlling factor (see LCR+ model).

LIMITING CORROSION RATE (LCR) MODEL

In the LCR model, it is assumed that the mass transport of (cathodic) corrosive species through the diffusion
boundary layer is the corrosion rate determining step. No limitation is set by charge transfer kinetics. This provides
a maximum or limiting corrosion rate. Therefore, the model is referred to as the Limiting Corrosion Rate (LCR)
model.

Corrosion is assumed to occur via the reduction of protons (H +), as in normal acid corrosion, and via the reduction
of carbonic acid (H2CO3)

H e 12 H 2
H 2CO3 e HCO3 21 H 2

Protons have to diffuse from the bulk region through the boundary layer to the wall, while the transport flux of
carbonic acid needs to reflect both diffusion of H2CO3 and hydration of CO2 in the boundary layer. Hydration of
carbon dioxide to carbonic acid in the diffusion boundary layer is an extra source of corrosive species to the
transport flux
k1
CO2 H 2O H 2CO3
k2

The hydration process is relatively slow, but there is a high abundance of carbon dioxide in the water phase.

Contribution of proton reduction to the corrosion rate

The maximum proton flux follows from Fick's law as

j k[ H ]bulk k 10 pH
where

D
k
D

is the mass transfer coefficient, with D the molecular diffusivity and D the depth of the diffusion boundary layer.
The depth of the boundary layer is about 50 m for a flow velocity of 1 m/s and a temperature of 25 C. The mass
transfer coefficient can be calculated from correlations for the Sherwood number, the standard dimensionless mass
transfer coefficient. For turbulent flow in a smooth pipe the following correlation holds (Treybal 1955)

kDH
Sh 0. 023 Re 0.83 Sc 1 / 3
D

where DH is the hydraulic diameter, defined as

4 x flow area
DH
wetted perimeter

For a rough surface, the Sherwood number needs to be replaced by one which contains the surface roughness as
a parameter. The Reynolds number, Re, and Schmidt number, Sc, are defined as

VDH
Re


Sc
D

respectively, where V is the (average) flow velocity and the kinematic viscosity of the aqueous solution.
Substitution leads to the following expressions for the mass transfer coefficient

V 0.83 D 0.67
k 0. 023
0.5 DH0.17

2
The relation between the mass transport flux of corrosive species j in kmol/(m s) and the corrosion rate in mm/y is

1 M Fe
CR j x1000x365x24x3600 = 1.13x108 j
2 Fe

2+
which applies to reduction reactions leading to the dissolution of half a ferrous ion (Fe ), where MFe is the iron
3
molecular weight (55.85 kg/kmol) and Fe the iron mass density (7800 kg/m ).

V 0.83 D 0.67 pH
CR 2. 6 x 10 6 10
0.5 DH0.17

Proton diffusivity and water kinematic viscosity are temperature-dependent factors in this relation. Collecting these
factors in one function, allows the limiting corrosion rate to be written as

V 0.83 ( pH 4 )
CR1 f1 (T ) 10
DH0.17

where f1 would be the corrosion rate at a flow velocity of V=1 m/s, a hydraulic diameter of D H= 1m, and pH=4.
Numerical values for f1 are shown in Figure 1. The last equation and Figure 1 allow limiting corrosion rates to be
calculated for any combination of temperature, pH, fluid velocity, and hydraulic diameter.
Figure 1. Contribution of proton and carbionic acid diffusion and carbon dioxide hyderation to limiting corrosion
rate for fluid velocity of 1 m/s, hydraulic diameter of 1 m, and pH=4.

Contribution of carbonic acid reduction to corrosion rate

The H2CO3 mass transport flux follows from the steady state material balance equation for carbonic acid, which
should include the hydration/dehydration reaction terms, and Fick's law. Assuming that CO2 will not deplete, in
view of its high abundance, the flux is found to be

D
j coth( ) [ H 2CO3 ]
D
where
D

r2

is the ratio of the diffusion and (dehydration) reaction boundary layer depths. Note that the flux equals classical
Fickian diffusion multiplied by the factor coth().

The depth of the (dehydration) reaction boundary layer is

r 2 D / k2

where k2 is the dehydration reaction rate. At 25 C, this depth is about 10 m. The depth of the diffusion boundary
layer follows from the Sherwood number and is about 30 m for a flow velocity of 1 m/s and a temperature of 25
C.

For D<r2 or <1 (flow-controlled corrosion) the transport flux is

D
j kD [ H 2CO3 ] [ H 2CO3 ]
D

while for r2<D or >1 (reaction-controlled corrosion)

D
j [ H 2CO3 ] k2 D [ H 2CO3 ]
r2
The carbonic acid bulk concentration can be calculated from the carbon dioxide concentration in solution and the
equilibrium ratio K for the hydration/dehydration reactions (K=k1/k2). The ratio K is almost constant over the
temperature range from 0 to 100 C (K~0.0026). The carbon dioxide concentration in solution follows from the
carbon dioxide partial pressure pCO2 (in bar) in the gaseous phase via Henry's constant H. This gives
[ H 2CO3 ] KHpCO2

The limiting corrosion rate can now be written as

D
CR 1. 13 x 10 8 coth( )KHpCO2
D

Collecting the temperature-dependent factors in one function gives under flow control

V 0.83 D 0.67 V 0.83


CR2 2. 6 x 10 6 0.5 0.17
KHpCO2 f2 (T ) 0.17 pCO2
DH DH

while under reaction control

CR3 1. 13 x 10 8 Dk2 KHpCO2 f3 (T ) pCO2

which latter corrosion rate is flow-independent. The f1 and f2 functions in the above equations, which correspond to
the limiting corrosion rates at pCO2=1 bar, V=1 m/s, and DH=1 m, are shown in Figure 1.

The temperature dependence of the carbonic acid diffusion contribution to the corrosion rate is only weak, which is
caused by cancelling temperature influences in the carbonic acid diffusivity, water viscosity, and CO2 solubility.

The total limiting corrosion rate is obtained after summing the contributions by protons and carbonic acid. At
relatively high CO2 pressure and high pH, corrosion proceeds via direct hydration of CO2 and there is little flow
dependency for flow velocities in the range between 0.1 and 1 m/s. At lower CO2 pressure and lower pH, proton
reduction is dominant, the rate of which is controlled by mass transfer, and there is a far stronger flow dependence.

Comparison with "de Waard-Milliams"

The LCR model gives corrosion rates in reasonable agreement with the model of de Waard and Milliams for flow
velocities around 1 m/s, which roughly mimics the mass transport situation of their autoclave tests. The original
nomogram model follows more or less the hydration reaction contribution to the corrosion rate, which can be
understood because the pH in the autoclave tests corresponded probably to saturation or super-saturation
conditions (i.e. above pH=5) and the proton diffusion contribution was therefore small (<10%). At 1 m/s also the
carbonic acid diffusion contribution is not very large (30 %). Hence, the original nomogram is based on
predominantly hydration reaction controlled tests. For higher velocities the diffusion or mass transfer contributions
increase and the two models will diverge. The LCR model also agrees with the updated deWaard-Milliams model
(NACE/91) for pH's below saturation (V~1 m/s).

COMPARISON WITH EXPERIMENTAL DATA

Comparison with the IFE KSC-II corrosion data confirms the model to provide an upper limit. For the 3.1 m/s tests
(unbuffered), there is good agreement. As expected, for higher velocities the LCR model overpredicts the corrosion
rate as no charge transfer limitation is included.

For the buffered tests (increased pH due to ferrous ions from corrosion) there are cases where the LCR model
underpredicts the corrosion rate. The most likely explanation is the extra corrosion caused by the presence of a
carbide network. Modern pipeline steels contain relatively little carbon (<0.1%) and an enhancement of the corrosion
rate on top of the LCR value due to the presence of carbide networks needs not to be considered. The same applies
to quenched and tempered tubing steels, which have no pearlite network. Consideration of network effects might be
required though for piping when using more conventional steels (e.g. ferritic/pearlitic St52 steel with say 0.2 %
carbon).

7.1.5. Corrosion product scaling

The scale factor accounts for the reduction of the corrosion rate due to the presence of
corrosion product film. In CO2 corrosion the main solid product is iron carbonate, FeCO 3. The scale factor is a function
of temperature and CO2 partial pressure (actually fugacity) and becomes significant only at temperatures above 60 C.

Formation waters can have a beneficial effect but also a detrimental effect on scale formation. For example a CaCO 3
scale can be protective as long as no FeCO3 scale forms, while for conditions where normally a protective FeCO 3
2+
scale would form, the presence of Ca may lead to a mixture film of calcium and iron carbonates with poor protective
properties. Also the role of chlorides on scale protective properties is not clear. For above reasons, the scale factor is
not applied when formation water is present (applies to all three corrosion rate models).

The scale factor is also not applied when the mixture velocity exceeds 20 m/s, assuming this velocity marks the onset
of erosion damage to the scale.

In the 1991 NACE model and the 1995 IFE fit model, the scale factor is applied irrespective of the iron carbonate
scaling tendency [7] it is even applied for iron undersaturation. However, there is sufficient experimental evidence that
iron undersaturation can prohibit scale formation [8]. For this reason the scale factor is not applied in combination with
both the LCR+ and LCR+99 models when the scaling tendency is too low as will be explained next.

The scaling tendency is defined as the ratio of iron carbonate precipitation rate and corrosion rate.
Precipitation rate
ST
Corrosion rate
It is assumed that when this ratio is below a critical value ST=0.5, any scale forming will be undermined by corrosion
and no full scale coverage of the steel surface is obtained.
The rationale behind the value ST=0.5 is that the precipitation rate and corrosion rate are defined on a mole basis.
Taking into account the volume difference between iron carbonate and iron of about a factor of 3, at a scaling
tendency of ST=0.5, the surface areas of a steel atom and an iron carbonate molecule are about the same. Computer
simulations, in which the processes of steel loss by corrosion and surface coverage by iron carbonate are dynamically
modelled, confirm the value of ST=0.5 to be a reasonable choice (see Appendix 5). In experiments by Van Hunnik et
al. [7] actual critical scaling tendencies were found to lie between 0.4 and 0.95 (depending on carbon content,
microstructure, and flow velocity).
Conditions which favour a high scaling tendency are: high temperature, low flow, high bicarbonate level (i.e. high pH),
2+
and high Fe level. Iron supersaturation must prevail to achieve ST>0.

In the previous version of HYDROCOR, the scaling tendency was calculated from the water bulk properties. In the
current version, the scaling tendency is based on the wall properties.
Organic acids are another factor in scale formation but are not necessarily related to formation water. Organic acids
are carried via the gas phase and will dissolve in condensation water, thus leading to more aggressive water than with
CO2 only. The effect of organic acids on the pH and iron solubility is included.

7.1.5.1. FeCO3 scaling computer simulation

Computer simulation of FeCO3 scaling

The dynamic process of precipitation of corrosion product and its competition with corrosion can be easily simulated
and visualised with the aid of a computer program written in Visual Basic.

Starting with an un-corroded piece of steel, corrosion progress can be simulated by taking out "grains" randomly from
the exposed steel surface not covered by iron carbonate scale. The iron carbonate precipitation process can be
simulated by dropping at random locations of the surface, carbonate "crystals" (proportional to the scaling tendency,
taking into account the surface area differences between steel and iron carbonate). The only parameters in the model
are the scaling tendency and time. The scaling tendency is the ratio of precipitation rate and corrosion rate (on mole
basis).
Examples of simulation snap shots are shown in Figure 4-1 for a number of scaling tendencies between 0 and 0.7. For
scaling tendencies up to ST=0.3, scale formation is undermined by corrosion and no full scale coverage is ever
obtained. At ST=0.5, undermining of the scale is no longer possible and a protective scale forms. Small pits may form,
even above ST=0.5, but they stop growing after a while.
Steel ST=0 ST=0.4

FeCO3 ST=0.1 ST=0.5

ST=0.2 ST=0.6

ST=0.3 ST=0.7
Figure 4-1
Computer simulation snap shots of steel corrosion and iron carbonate precipitation
(blue = corrosive environment, grey = steel, red = iron carbonate). For scaling tendencies above ST=0.5, scale
protection occurs

7.1.6. Oil wetting

Credit is taken for the protection by an oil phase, the protection depending on water cut and flow conditions.

First it is established whether the liquid flow velocity is high enough to avoid water drop-out. The critical velocity for
water drop-out is taken at 1.5 m/s, based on field measurements in 1994 in a number of PDO oil pipelines. This value
replaces the previous "Wicks and Fraser" value of 1 m/s. Actual critical velocity for water drop-out is oil specific and
the above rule may be very conservative. Lab and/or field measurements are required to establish the actual water
distribution, including the tendency to form a stable emulsion. This information cannot be obtained from a simple
model.

For liquid flow velocities above 1.5 m/s, it is assumed than a water-in-oil dispersion exists for water cuts below 40%
and there will be zero corrosion: Foil=0. For higher water cuts, water wetting and corrosion are assumed proportional to
the water cut: Foil=water cut/100%.
For annular dispersed gas/liquid flow, it is assumed that the water phase can always wet the pipe wall, irrespective of
the above conditions (Foil =1). This is a very conservative and perhaps questionable assumption, but three-phase flow
tests would be required to find out under what conditions oil wetting occurs.

For stratified gas/liquid flow, the liquid flow velocity is the velocity of the water/oil layer at the bottom. For gas/liquid
slug flow, the liquid film velocity between the slugs is taken.
For condensates (NGL's ) inhibiting effects are minimal and are ignored.

7.1.7. Alcohol injection

The mitigating effect of glycol or methanol within the aqueous phase is included utilising empirical correlations for the
inhibiting effects of MEG, DEG, TEG, or methanol. The alcohol factor is a function of the alcohol type and alcohol
fraction in the water.

The partitioning of the alcohol over the phases (bottom water, vapour phase, and condensing water) in a long pipeline
is quite complex and is described by a separate model.
7.1.8. Effect of total system pressure

The CO2 fugacity, fCO2, rather than the CO2 partial pressure, pCO2, is utilised in the corrosion rate equations, this to
account for the non-ideal, reduced CO2 activity at high pressure. The fugacity coefficient is used from the 1991 NACE
paper by the Waard et al. [2]. At 200 bar and 25 C, the fugacity is a factor of about two smaller than the partial
pressure.

7.1.9. Corrosion inhibitor injection

Corrosion control via corrosion inhibitors is evaluated via the equation for the average corrosion rate over the inhibited
and non-inhibited periods

CR f x CRi (1 f ) x CRu

where f is the fraction of time the inhibitor is working (availability), CR i the inhibited corrosion rate, and CRu the un-
inhibited corrosion rate as obtained from the HYDROCOR corrosion rate models. Persistency effects are ignored in
the above equation. A default upper limit inhibited rate of 0.2 mm/y may be adopted for Cr i [9]. Extensive laboratory
testing indicates that this default inhibited rate of 0.2 mm/y is conservative. The CO 2 corrosion position papers [17-18]
therefore recommend a lower residual corrosion rate of 0.1 mm/y (up to 120 C). Consideration must be given to the
adoption of even lower values based on representative corrosion inhibition testing.

The inhibited corrosion rate cannot exceed CRu.

HYDROCOR is still assuming the inhibitor fails above a mixture velocity of 20 m/s. Actually, failure of an inhibitor due
to high velocity (high shear stress) depends on the inhibitor concentration. Higher concentrations afford a higher
critical velocity, which can well exceed the "old" 20 m/s velocity limit.

7.2. Organic acid corrosion

Whilst normally more than one fatty acid is present, acetic acid (ethanoic acid) is taken as the base for the corrosion
rate prediction. Often, acetic acid has the highest concentration.

Electrochemical reduction of the following two species are considered for the calculation of the corrosion rate:

+
protons (H )

un-dissociated acetic acid (HAc)

Calculation of the corresponding corrosion rates is based on assuming mass transfer and charge transfer are the main
rate-determining factors. Writing the limiting corrosion rates from these two mechanisms as

CRmt=kC

(-E/T)
CRct=Ae

the corrosion rate follows from (series resistance model)

1/CR=1/CRmt+1/CRct
+
where k is the mass transfer coefficient, C the concentration ([H ] or [HAc]), and T the absolute temperature.

If only the contribution from the acidity (protons) would be taken, failures of carbon-steel facilities by organic acid
corrosion cannot be explained. With the additional reduction of un-dissociated organic acid field failures can be
explained (e.g. Barendrecht water drain pipe failure, 1998).

The mass transfer coefficient follows from chemical engineering equations, the species concentrations from the
dissociation equilibrium.
For the charge transfer a limiting corrosion rate of 10 mm/y at 20 C is assumed, with a doubling of this rate for a 20
C temperature increase.

7.3. H2S corrosion

When CO2 is present in combination with H2S, three corrosion regimes can be distinguished:

pCO2/pH2S>500: CO2 corrosion alone

pCO2/pH2S between 20 and 500: slightly sour environment. In this regime localised corrosion can occur, but the
corrosion rate was found not to exceed the CO2 corrosion rate without H2S.

pCO2/pH2S <20: H2S corrosion dominates. There is an enhanced pitting risk, particulalry for conditions with free
sulfur and chlorides.

In the sour corrosion regime (pCO2/pH2S<20), field corrosion rates as high as 20 to 30 mm/y have been seen in our
fields, notably in Canada. Prompted by these high corrosion rates, a start was made with the development of a sour
corrosion model. For its development, tests are being carried out in autoclaves (~200 ml) at various levels of CO 2 and
H2S (1 to 10 bar), various levels of chlorides (1000 to 100000 ppm), with and without elemental sulfur. While a
magnetic stirrer is used to generate flow, the flow velocity is limited to very low values (<1 cm/s) to avoid the sulfur
starting to float around and losing contact with the coupon. The low flow is also believed to simulate worst case
conditions. Based on the experiments carried out so far, the following preliminary model was proposed:

CR F p CR sweet

where Fp is the pitting factor. The pitting factor is a function of elemental sulfur presence and chloride level in the
water. As oxygen can lead to the production of free, elemental sulfur in sour service, presence of oxygen is interpreted
as having elemental sulfur present.

In the sour regime, the various factors and effects required re-consideration. An overview of our current suggestions is
given below

Top-Of-Line corrosion

Currently no TOL corrosion model is available for the sour regime. Use of the sweet corrosion model is seen as a
conservative approach, assuming there are no agents that can destroy the TOL iron sulfide scale.

Scale factor
The slowing of corrosion by iron sulfide scale poses the greatest challenge in sour corrosion modeling. Formation of
an iron sulfide scale is one of the main characteristics of sour corrosion, even at relatively low temperatures and
explains the low field corrosion rates often seen. Corrosion in the field relates to failure of the iron sulfide scale, rather
than the absence of scale. However, no method is currently available to distinguish between protective and non-
protective situations. Obviously, the CO2 corrosion scale factor for iron carbonate is of no use in the sour corrosion
regime.

Comparison with data

The SOGACOR report shows a comparison of the sour corrosion model with our lab data. Similar and worse results
were obtained with other (commercial) sour corrosion models demonstrating that considerable further work is required
before an acceptable sour corrosion model will be available. The current models are far too conservative. The scatter
is explained by differences in scale protection. The protection will differ between tests, even for the same test
conditions. In most tests, the scale will slow corrosion to levels well below the bare-steel (CO 2) corrosion rate. Only for
cases with a strong disturbance of the scale by free sulfur and/or chlorides, will corrosion rates be close to or higher
than the bare-steel (CO2) corrosion rates.

Another shortcoming of the current sour corrosion modeling is that the sweet corrosion rate is taken as the reference
whereas H2S and free sulfur are both corrosive towards carbon steel without requiring other corrodents. This issue
needs further attention. The current model does not work for systems containing no CO 2.

7.4. Bacterial corrosion

In the early 1990's, we began to develop an approach to assist ranking of oil transport pipelines for susceptibility to
micro-biologically influenced corrosion (MIC). The outcome of this work was a "predictive flowchart" where the
susceptibility of a pipeline could be assessed on the basis of details of water chemistry and pipeline operational
parameters.

MIC of mild steel manifests as a localized, pitting form of attack, following the development of a surface biofilm. The
anaerobic environments of oil transport pipelines often support the growth of biofilms, which almost invariably contain
sulfate-reducing bacteria (SRB), a major cause of MIC. SRB-containing biofilms generate H 2S, which precipitates iron
sulfide within the biofilm. These sulfides are cathodic to the steel and may greatly exacerbate corrosion at anodic
sites.

The physico-chemical and biological interactions, which take place between the biofilm and the environment in a mild
steel pipeline, are very complex. They may be interdependent on other processes not directly related to biological
activity. This makes modeling the suite of corrosion phenomena that constitute MIC an extremely difficult task.
Prediction of pipe wall penetration rates attributed to MIC-induced localized corrosion is unreliable, because of the
uncertainty of the onset of pitting and whether it proceeds at a constant rate. Nevertheless, the original predictive flow
chart was translated into a MIC corrosion rate predictor, realizing that corrosion rates can be uncertain. Corrosion rate
values are very approximate. The main idea is to link the risk of MIC and the influencing parameters.

Predicting the susceptibility of a pipeline to MIC is facilitated by considering 1) whether the pipeline environment can
support microbial activity and biofilm formation, and 2) how pipeline operational parameters affect the microbiology.
The premise is that if an active biofilm can form inside a pipeline, then that line may suffer from biofilm-associated
corrosion.

The MIC corrosion rate calculation is based on the following equation

CR C x F p

with

F f 1 x f 2 x....... f n

where C is a constant (C=2 mm/y), the f's are factors for the various influencing parameters, and p is a power law
index (0.57). An overview of the factor values is given in Table 4.

Influencing parameters

The following parameters influence MIC:



Presence of water: if the crude and pipeline are free of water, there will be no MIC.

Water wetting: MIC is only to be considered at separated flow conditions; if all water is entrained in the oil, the risk of
MIC is significantly reduced.

pH of water: SRB growth is largely confined between pH's of 5 and 9.5.

Salinity or total dissolved solids (TDS): most SRB will grow best at < 60 g/l TDS.

SRB may still grow if TDS > 60 g/l, but this must be confirmed by growth in a nutrient medium with the same TDS
concentration.
o
Temperature: different types of SRB can grow at temperatures between 4 and 110 C.
o
If the temperature is above 45 C, it is necessary to confirm that SRB isolated from the pipeline can grow at the
pipeline temperature.

Microbial activity is generally limited by extremes of these parameters. Although certain microorganisms can grow
outside the normal limits, they are usually confined to extreme habitats. The limits for pH and temperature are well
established for bacteria that could occur in oil transport pipelines, but tolerances for salinity are less well understood.
The important consideration with these parameters is whether bacteria isolated from a system can grow under the
prevailing conditions. For example, in a pipeline carrying high salinity produced fluids, it may be possible to isolate
bacteria using standard isolation media, but these media may have a considerably lower salinity than the water phase
in the pipeline. There may be no significant bacterial growth at pipeline conditions and therefore no significant impact
with respect to MIC.
Demonstration of growth at pipeline conditions is therefore an essential part of the assessment of susceptibility to
MIC.

Impact of nutrient status of water phase

The following nutrients have an influence on the corrosion rate:

Sulfate ion concentration: SRB growth is severely reduced at sulfate levels < 10 mg/l unless an alternative electron
acceptor is present.

Total carbon (C) from fatty acids such as formate, acetate, propionate, & butyrate: SRB growth is severely restricted
if utilizable carbon is < 20 mg/l.

Nitrogen (as utilizable N): SRB growth is severely restricted if utilizable nitrogen is less than 5 mg/l

Critical C:N ratio: SRB growth is the most prominent if C:N <10.

Nutrient availability directly influences microbial activity and biofilm formation both qualitatively and quantitatively.
Carbon and nitrogen are the two of the most important nutrients, which can be readily quantified. Sulfate is also
important for SRB. It is the principal respiratory electron acceptor, which can ultimately have a significant influence on
MIC. There are threshold concentrations below which these nutrients cannot be effectively utilized by microorganisms.

When one of the nutrients essential for microbial activity falls below the threshold, this activity ceases until more
nutrients become available. It should be noted that a water analysis indicating a nutrient deficiency does not mean that
no MIC can take place, if previously nutrients were available. MIC associated with a biofilm that was formed when
nutrients were available, but which subsequently faces a nutrient shortage, may continue for some time.

Additionally, the ratio of carbon to nitrogen is important. A ratio of carbon to nitrogen of approximately 10:1 is required
to sustain microbial growth. The ratio will indicate the limiting nutrient, useful information when looking at options for
controlling microbial activity (e.g. a production chemical could change the ratio in such a way as to encourage
microbial activity).

Some bacteria can survive and even grow in the complete absence of sulfate, they simply use an alternative electron
acceptor, such as nitrate. This concept has been utilized to prevent reservoir souring from water floods. However, this
form of metabolism is unlikely to be significant in oil transport pipelines. Apart from sulfate, other oxidized sulfur
species may be utilized, such as thio-sulfate or bi-sulfite. Hence if there are sources of these species present in the
water, then they may represent an important increase in the potential for sulfide generation if the level of sulfate is very
low.

Note that the threshold values given for nutrients are based on rather few data. These values should be substantiated
with more experimental data to obtain a higher level of confidence since oil field waters often contain key nutrients
close to threshold limits.
Impact of flow rate

Flow rate will directly influence the nature of biofilm formation and the rate of nutrient delivery. As flow rate increases,
biofilms become less bulky and more adherent. However, above a certain threshold, the initiation of biofilm formation
is significantly limited. The transition zone is considered to be between 2 and 3 m/s. At the other extreme, stagnation is
often associated with the severest MIC incidents. A pipeline's flow history could be important in determining current
status of the line and subsequent risk of MIC.

Impact of debris on the bottom of the pipeline

Presence of debris exacerbates MIC, because the presence of debris can create stagnant conditions in the bottom of
low velocity lines, where biofilms can be established.

Impact of pigging frequency

Pigging can potentially have a significant effect on biofilm and MIC. The more frequent the pig runs, the less time
biofilm has to recover. A pigging frequency of at least once every two weeks will be very effective in controlling biofilm.
No detailed data are available on different pigging intervals versus biofilm recovery rates or MIC, although there is
some published experience of the impact of pigging as a treatment for MIC. The extent to which pigging is effective in
controlling biofilm-associated corrosion depends on how well a corrosion process was established prior to the pigging
program. Since pigging has a significant impact on the occurrence of MIC, the frequency should take a prominent
place in the prediction module.

Impact of oxygen ingress

The highest biofilm-associated corrosion rates are seen when oxygen periodically accesses the sulfide-rich, anaerobic
region within SRB biofilms.

Impact of a biocide treatment

Biocides are used to control unwanted microbial activity. If a biocide is used in a pipeline, the risk of a failure due to
MIC should be significantly reduced, or even eliminated. It is essential that the biocide is selected carefully and the
treatment regime is appropriate and applied properly and consistently. For practical purposes, the inclusion of biocides
as a parameter in a predictive scheme would need to be based on the assumption that the correct biocide has been
selected and applied in such a way that it can function as intended. With this in mind, it is possible to make some
simple rules regarding the type of biocide and the application dose and frequency for different types of water
chemistries, watercuts etc.

Impact of operational history of pipeline


The following operational history parameters have an influence on MIC:

Operational age of pipeline in years

Total operational downtime:

Long periods of stagnation promote MIC.

Other parameters

Other parameters may also play a role in MIC. So far, these parameters have not been included in the model.

An example is production chemicals. Production chemicals can impact MIC in a number of ways. They may be a
source of nutrients to stimulate microbial activity. They may interact with and reduce the efficacy of biocides being
used to treat a bio-fouled system. Corrosion inhibitors could act to reduce MIC in some circumstances. Previous
studies have shown that production chemicals can indeed affect the performance of some biocides, and vice versa.
Phosphorus and nitrogen-containing chemicals have been shown to act as nutrient sources. Since production
chemicals are often used continuously in oil transport pipelines, this parameter needs to be addressed in a future
predictive module.
7.5. Oxygen corrosion

The current oxygen corrosion model is based on assuming that the rate-determining step is the diffusion of oxygen
over the diffusion boundary layer. Standard Sherwood mass transfer calculations are carried out to calculate the
corrosion rate.
8. Other essential models in HYDROCOR

8.1. Water chemistry

The calculation of the pH, needed for the bare-steel corrosion rate, follows from the water composition and aqueous
chemistry principles. The parameters are the CO2 partial pressure, bicarbonate level, dissolved iron concentration,
and temperature. The development of the dissolved iron level along the line is calculated from an iron mass balance; it
includes iron carbonate precipitation and allows for iron supersaturation

8.2. Alcohol/water distribution

The alcohol/water distribution (partitioning) over the aqueous phase and vapour phase along the pipeline is calculated
[1 ]
using a model by Lammers 0 . The model uses phase equilibrium and interfacial mass transfer principles.

8.3. Gas/liquid two-phase flow

Corrosion rate calculation for two-phase flow conditions requires details such as flow pattern, liquid hold-up, actual
liquid flow velocities, liquid hydraulic diameters, length of film and slug zones, etc. Two-phase flow details in
HYDROCOR are obtained from calling a Shell proprietary model, the so-called KSLA method [18].

The KSLA method distinguishes the following flow patterns: stratified (wavy), slug (or intermittent), annular dispersed,
and (dispersed) bubble. For the separated flow patterns (stratified and annular dispersed), the actual liquid velocity is
taken as the relevant flow velocity for the corrosion calculations, which is the liquid film velocity for the stratified and
annular dispersed flow patterns. The hydraulic diameter of the liquid phase is taken as the relevant diameter, see [Ref.
6] for details.

Stratified wavy Annular dispersed

Slug (intermittent) Dispersed bubble


Flow patterns for (horizontal) two-phase flow

For slug flow, the flow pattern is divided into a slug part and a film part. Corrosion rates are separately calculated for
these parts. For the slug zone, the mixture velocity, which is the sum of the liquid and gas superficial velocities, is
taken as the relevant flow velocity. The hydraulic diameter equals the pipe diameter. For the film zone the same
procedure applies as for the separated flow patterns. The overall corrosion rate is calculated from averaging over the
two zones.
For the two-phase flow calculations, free water and liquid hydrocarbons are lumped to one liquid phase. Physical
properties follow from averaging over the volume fractions.

8.4. Pressure and temperature profiles

The pressure gradient is automatically generated from the two-phase flow model. The temperature profile is obtained
from a heat transfer model.
2
The user has to provide the pipe heat transfer coefficient pipe. Default value is 20 W/m K, which applies to a "typical"
concrete weight-coated offshore pipeline (see table below). The thermal resistance in this case is mainly determined
2
by the concrete. Without concrete cover, the coefficient would be 100 W/m K. For a bare offshore pipeline, the
2
coefficient would be typically 2000 W/m K. The heat transfer coefficient is rather insensitive to the absolute pipe
2
diameter. A typical value for an insulated (50 mm foamed polypropylene) offshore pipeline is 6 W/m K unburied and 4
2
W/m K under 1 m burial.

Pipe heat transfer coefficient for "typical" insulated offshore pipeline

8.5. Heat transfer model

The temperature profile is calculated from the heat transfer from the pipeline to the outside environment. The heat
balance is written as:

dT
( C p Q ) i (T Ta )D (1)
dx
3
where (kg/m ) is the mass density, Cp the heat capacity at constant pressure (J/kgK), Q the volumetric flow rate
3 2
(m /s), T the temperature (K), the effective heat transfer coefficient (W/m K), Ta the outside environment
temperature (K), and D the inside pipe diameter. The summation is over the vapour, condensate and aqueous phases.
The same temperature is assumed for the phases.

The solution of Equation (1) for a pipe section with length l is:

Tout Ta D
exp l (2)
Tin Ta ( C Q )
p i

where the subscripts in and out refer to the values at the inlet and outlet of the section.
The heat transfer coefficient is given by:

1 1 1
(3)
convective pipe

where convective is the convective heat transfer coefficient and pipe the total pipewall heat transfer coefficient, which is
determined by the thickness and heat conductivity of the following layers: pipeline steel, coating and concrete:

1 1 1 1
(4)
pipe steel coating concrete

Each layer coefficient is given by


2
(5)
Douter ln( Douter / Dinner )

where is the thermal conductivity (W/mK), Douter the outer diameter (m) and Dinner the inner diameter (m).

2
The user of HYDROCOR has to provide the pipe heat transfer coefficient pipe. Default value is 20 W/m K
The advantage of having the user to provide the pipe heat transfer coefficient is that only one input parameter is
required. Otherwise, at least the thickness and conductivity of each thermal layer would have to be provided. In most
cases it will not be easy to get accurate data of the various thermal layers anyway. Therefore, it is suggested to use
the above information as a guideline to prepare a reasonable guess for the pipe heat transfer coefficient and to
examine the sensitivity of the corrosion rate.

The convective coefficient is calculated internally in the program from:


(6)
DH / Nu
with the Nusselt number correlation:

Nu 0.023 Re 0.83 Pr1/ 3 (7)

where Re is the Reynolds number and Pr the Prandtl number:

vD H
Re (8)

Cp
Pr (9)

2
with v the actual phase velocity (m/s), DH the hydraulic diameter (m), and the viscosity (Ns/m ). A convective heat
transfer coefficient is calculated for each of the phases. The overall coefficient follows from averaging over the phases
according to their "wetting" of the pipewall.
The heat transfer coefficient between concrete and the environment is not considered in the above formulation, but
can be incorporated in the pipe heat transfer coefficient by the user if this would be required.
9. MULTCASE examples

MULTCASE contains a number of example cases, two of which are considered in this manual. Users may want to look
not only for the corrosion rate but also for the following:

Flow pattern and other flow details (such as liquid hold-up, relative slug length, film velocity, wetted perimeter,
etc.) This helps to get an idea about where the corrosion should be expected, which can help the inspection
engineer in intelligent pig surveys. In the Troll example below the flow pattern is stratified wavy. The hold-up is
0.3 %. So there is only a relatively small aqueous phase stream at the bottom.
Scaling tendency. This helps to get an idea about the type of attack. If the scaling tendency is close to one, a
protective iron carbonate layer may develop. Corrosion is expected to develop in the form of localised attack
rather than uniform attack of the wetted perimeter.

Troll 36" wet gas pipelines


.
HYDROCOR input/output screen for the Troll 36" wet gas pipelines Glycol is applied to reduce the corrosion rate from
about 2 mm/y to about 0.2 mm/y, which rate is judged acceptable with 10 mm corrosion allowance for 50 years project
lifetime.

Corrosion damage case

The second example refers to an actual corrosion damage case of a 10" high pressure gas pipeline, which
corresponded with a corrosion rate of 0.7 mm/y.

The calculated corrosion rate for the example is about 2.5 mm/y, whereas the field corrosion rate was 0.7 mm/y. This
confirms HYDROCOR to predict corrosion conservatively for this particular case. The order of magnitude is correctly
predicted. Obviously, the actual corrosion is far too severe for a pipeline with a wall thickness of only 4 to 5 mm.

The flow pattern is stratified way. The hold-up is less than 0.2%. The wetted perimeter at the bottom is therefore only a
few cm's and corrosion is expected at the bottom only. The scaling tendency reaches values up to about 0.6.
Therefore, scales may be present in the line and damage should be localised rather than uniform over the wetted
perimeter. This is confirmed by the case. Grooving corrosion at the bottom appeared to the cause of the pipeline
rupture.
10. Changes between versions

10.1. WETGAS7.2/FLOWLINE7 & HYDROCOR1.04

The following list summarises the main changes between WETGAS7.2 /FLOLINE7 and HYDROCOR1.04.

Single-point analysis is possible, which makes FLOLINE7 redundant.

More extensive explanations are given in the Note or Comment boxes. A problem is that EXCEL7 keeps the boxes
much too small to read all the text. In EXCEL8 the boxes can be enlarged (but they default again to small boxes
when saved as an EXCEL7 file). Excel 9 that is part of Office 2000 has the same attributes as Excel 8 in this
respect. Printing of the notes is an alternative.

A distinction between oil and NGL's is made but the user has to decide between the two (the input cell only affects
the oil factor).

Based on 1997 R&D work on scaling, it was decided to let the scale factor of de Waard only be in effect if the
scaling tendency, ST, based on the bulk properties of the aqueous phase, exceeds a certain critical value (ST=0.5).

For TOL (top-of-line) corrosion prediction we have incorporated model B discussed in a recent OP report [18]. This
will generally lead to lower TOL corrosion rates than before, except for temperatures above 80 C.

Imperial units can be utilised in the sheet "Main". Automatic conversion to the right input values is taken car

10.2. HYDROCOR1.04 & HYDROCOR2.01

Major changes:

Less conservative version of the LCR+ model was implemented, referred to as LCR+99 model.
For the assessment of the occurrence of (protective) FeCO 3 scale formation, evaluation of the scaling tendency
is now based on the wall chemistry rather than the bulk chemistry.
For Top-Of-Line (TOL) corrosion under stratified flow conditions, the iron supersaturation model (model D) was
implemented.
Corrosion caused by organic acids was added. Corrosion by organic acids (simulated by lumping them to acetic
acid) is assumed to originate from (a) their acidity and (b) the direct reduction of the acetic acid (HAc) species.
+
For both these contributions to the corrosion rate, it is assumed that mass transfer of the H ions and HAc
species is the rate-determining factor.
An entry was created for the amount of dissolved CO2 in water to better cope with liquid-full systems (no free
gas).

Minor changes:

A new node-generator was implemented, which automatically generates shorter section lengths at the inlet of the
pipeline.
The temperature (axial) gradient model was improved.
Temperature dependency was included in the NGL/oil viscosity (Lewis-Squires model, see Reid et al.).
Cases can be directly loaded/stored from MULTCASE.

10.3. HYDROCOR2.01 & HYDROCOR3.01

Major changes:

Shell Canada's model for sour corrosion, SOGACOR, was included.

The user manual was made available as a HTML Help file.

Minor changes:

Free elemental sulfur and chlorides are new input parameters in the spreadsheet "Main".
11. References

1. M.B. Kermani and L.M. Smith, "CO2 corrosion control in oil and gas production. Design considerations",
European Federation of Corrosion Publications"
2. C. de Waard, U. Lotz, and D.E. Milliams, "Predictive model for CO 2 corrosion engineering in wet natural gas
pipelines", CORROSION/91, paper no. 577, (Houston, TX: NACE International, 1991)
3. A. Dugstad and K. Videm, "Kjeller Sweet Corrosion-II. Final report", report IFE/KR/F-90/008 KSC-II-59, Institutt
For Energiteknikk, Norway, 1990
4. C. de Waard and U. Lotz, "Prediction of CO2 corrosion of carbon steel", CORROSION/93, paper no. 69,
(Houston, TX: NACE International, 1993)
5. C. de Waard, U. Lotz, and A. Dugstad, "Influence of liquid flow velocity on CO 2 corrosion: a semi-empirical
model", CORROSION/95, paper no. 128, (Houston, TX: NACE International, 1995)
6. B.F.M. Pots, "Mechanistic models for the prediction of CO 2 corrosion rates under multi-phase flow conditions",
CORROSION/95, paper 137, (Houston, TX: NACE International, 1995)
7. E.W.J. van Hunnik, B.F.M. Pots, and E.L.J.A. Hendriksen, "The formation of protective FeCO3 corrosion product
layers in CO2 corrosion", CORROSION/96, paper no. 6, (Houston, TX: NACE International, 1996)
8. A. Dugstad, "Corrosion tests at high temperature and high CO 2 pressure", report IFE/KR/F-149, Institutt For
Energiteknikk, Norway, 1997
9. K.N. Ho and S.D. Kapusta, "Experience with carbon steel pipelines for wet gas transportation", Materials and
Corrosion Specialists Meeting 1996, paper F7
10. J.N.L.L. Lammers, "Phase behaviour of glycol in gas pipeline calculated", Oil and Gas Journal, April 15, 1991, p.
50
11. B.F.M. Pots and R.A. Connell, "Manual for HYDROCOR spreadsheet for the prediction of CO 2 corrosion in multi-
phase pipelines transporting wet hydrocarbons", OP.98.20493.
12. B.F.M. Pots and E.L.J.A. Hendriksen, TOL corrosion, OP.98.20046
13. A. Dugstad et al., "Kjeller Sweet Corrosion III - Final report", report IFE/KR/F-93/091, Institutt For Energiteknikk,
Norway, 1994
14. R. Nyborg et al., "Kjeller Sweet Corrosion V - Final report", report IFE/KR/F-98/075, Institutt For Energiteknikk,
Norway, 1998
15. J.H. Gerretsen, "CO2 corrosion Results of EMR project KSC-II at IFE, Norway An in-house spot-check",
ARGR.93.123
16. R. Nyborg, "CO2 corrosion in carbon steel with protective corrosion films in flowing water with salt Final report
KSC-IV", report IFE/KR/F-95/032, Institutt For Energiteknikk, Norway, 1995
17. B.F.M. Pots, S.D. Kapusta and R.A. Connell, "Position paper on assessing the feasibility of carbon steel in CO 2
service", SIEP 98-5565
18. B.F.M. Pots, S.D. Kapusta, R.A. Connell, and I.J. Rippon, "Assessing the feasibility of carbon steel in corrosive
service", SIEP 99-5661
19. B.F.M. Pots, R.V.A. Oliemans, and N. Trompe, "The KSLA method for gas/liquid two-phase pipe flow calculations.
Summary of physical models and empirical correlations and testing with laboratory and field data",
AMGR.88.3154
20. C. de Waard and D.E. Milliams, Corrosion, Vol. 31 (1975): p.177
21. R.E. Treybal, "Mass Transfer Operations", McGraw-Hill, 1955
22. D.A. Palmer and R. van Eldik, Chem. Rev., 83 (1983): pp. 651-731
23. S. Turgoose, R.A. Cottis, and K. Lawson, ASTM Symposium on Computer Modelling for Corrosion, San Antonio,
Texas, 12-13 November 1990
24. S. Nesic, "Progress in the understanding and prevention of corrosion", p. 539, Eds. J.M. Costa and A.D. Mercer,
University Press, 1993
25. J.S. Newman, "Electrochemical systems", Prentice Hall, 1991
26. A.J. Bard and L.R. Faulkner, "Electrochemical methods: fundamentals and applications", Wiley, 1980
27. J.T. Davies, "Turbulence Phenomena", Academic Press, 1972
28. J.N. Butler, "Carbon dioxide equilibria and their applications", Addison-Wesley, 1982
29. L.G.S. Gray, B.G. Anderson, M.J. Danysh, and P.R. Tremaine, " Mechanisms of carbon steel corrosion in brines
containing dissolved carbon dioxide at pH 4", CORROSION/89, paper no. 464, (Houston, TX: NACE
International, 1989)
30. R.V.A. Oliemans, "Multiphase Science and Technology for Oil/Gas Production and Transport", Lecture notes for
Course P2.24 on Multiphase Flow in Wells and Pipelines: Shell Training Centre, Noordwijkerhout, The
Netherlands
31. J.W. Palmer, J.L. Dawson, and G.A. Davies, "Gas phase corrosion inhibition project. Summary report", CAPCIS
report with reference CJPBBX, December 1992
32. S. Nesic, "CO2 corrosion of carbon steel in two-phase flow. The database", IFE report no. IFE/KR/F-93/058,
February 1993 (PROFF project)
33. S.D. Kapusta, B.F.M. Pots, and R.A. Connell, "Corrosion management of wet gas pipelines", CORROSION/99,
paper no. 45, (Houston, TX: NACE International, 1999)
34. B.F.M. Pots and E.L.J.A. Hendriksen, "CO2 corrosion under scaling conditions - The special case of Top-Of-Line
corrosion in wet gas pipelines", CORROSION/2000, paper no. 31, (Houston, TX: NACE International, 2000)
35. M. Simon Thomas et al., "Deterministic pipeline integrity assessment to optimize corrosion control and reduce
cost", CORROSION/2002, paper no. 02075 (Houston, TX: NACE International, 2000)
36. B. Hedges, D. Paisley and R.Wollam, "The Corrosion Inhibitor Availability Model", CORROSION/2000, paper no.
34 (Houston, TX: NACE International, 2000)
37. I.J.Rippon, "Carbon Steel Pipeline Corrosion Engineering: Life Cycle Approach", CORROSION/2001, paper no.
55 (Houston, TX: NACE International, 2001)
38. I.J.Rippon and M.J.J.Simon Thomas, "Selection of Corrosion Control Options to Optimise Production Field
Development", CORROSION/2002, paper no. 02278 (Houston, TX: NACE International, 2002)

You might also like