You are on page 1of 102

Corrosion Prediction Modelling

A guide to the use of corrosion prediction models for risk assessment in oil and gas production and transportation facilities
A J McMahon, D M E Paisley
Sunbury Report No. ESR.96.ER.066 dated November 1997

Main CD Contents

Contents
Summary

Page

Acknowledgements

Introduction

"Cassandra 98" Corrosion Prediction Spreadsheet by A J McMahon Introduction Quick Start Limitations of Corrosion Prediction Models Detailed Description of the Spreadsheet Comparing Output from the "Cassandra 98" Model with Field Data Appendix 1: Henry's Law Constans for CO Dissolved in Brine 2

5 6 8 11 27 29

The Use of Corrosion Prediction Models During Design by D E Paisley Introduction 31 Important Factors not Covered by the Corrosion Model 35 Effect of Corrosion Inhibitors 42 Predicting the Effectiveness of Corrosion Inhibitors 48 'The Inhibitor Availability Model' Recommended Values for use in the Inhibitor Availability Model 51 Comparisons of the Inhibitor Availability Model with BP's Previous Model 62 Corrosion Rates of Low Alloy Steels 64 Preferential Weld Corrosion 65 Effects of Pitting 66 Choosing an Optimum Corrosion Allowance 67 Applying Models to Different Flow Regimes 69 Applying Models to Transportation Equipment 72 Applying Models to Process Equipment 86 Flow Velocities in Process Pipework 89 Economic Tools to Use During Materials Selection 92

References

95

Installation of the Cassandra 98 Excel Workbook

97

Summary

This document decribes BP's current approach to Corrosion Prediction and its use during the design of pipelines and facilities. It is divided into two sections. The first section introduces a new prediction spreadsheet called Cassandra 98* which is BP's implementation of the CO2 prediction models published by de Waard et al. It builds on these models to include BP's experience of such systems. The pocket inside the front cover of this report contains a floppy disc which contains the necessary programs and spreadsheets to run it together with a set of installation instructions. The second section discusses how the prediction model may be used for design purposes and it introduces several improvements from previous guidelines. These include the use of the probabilistic approach to corrosion prediction and the use of corrosion inhibitor availabilities instead of efficiencies. It also discusses the use of "corrosion risk categories" as a way of quantifying the corrosion risk at the design stage. The floppy disc also contains a spreadsheet for calculating the risk category. To illustrate the points made examples have been obtained from many BP assets worldwide. Where financial data are shown it is from 1997. Since this subject is continually changing it is anticipated that these guidelines will be updated in future years and so any comments or suggestions regarding either the content or appearance of them would be very welcome.

*In Greek mythology Cassandra was the daughter of Priam and Hecuba. She was endowed with the gift of prophecy but fated never to be believed. She is generally regarded as the prophet of disaster........especially when disregarded.

Acknowledgements

The authors would like to thank the following BP staff for their contributions to these guidelines. Jim Corbally Laurence Cowie Mike Fielder Don Harrop Bill Hedges Will McDonald Tracy Smith Simon Webster Richard Woollam

Introduction

Carbon dioxide corrosion represents the greatest risk to the integrity of carbon steel equipment in a production environment. Compared with the incidences of fatigue, erosion, stress corrosion cracking or overpressurisation, the incidences of CO2 related damage are far more common. Unfortunately, the engineering solutions to eradicating the CO2 corrosion risk require high capital investments in corrosion resistant materials. As Figure 1 shows, providing a corrosion allowance of 8 mm to carbon steel flowlines costs a significant sum at circa US$1 million per 5 km but even this is insignificant in terms of the costs of the various corrosion resistant flowline options.
Figure 1: Fully Installed Costs for Various Flowline Materials Options in Colombia (1997)
35

30

Duplex SS

25

20 Cost per 5 Km ($mill) 15

Bi-metal 13Cr liner

13%Cr 10

Carbon steel 8mm ca 0 6 8 10 12 14 16 18 20 22

Carbon steel, no ca 24 26 28 30

Nominal Flowline Diameter - Inches

Similar relative costs are incurred when specifying corrosion resistant materials downhole or in facilities. This is rarely justified. For this reason, CO2 corrosion of carbon steel will always be a problem that BPX has to deal with. Managing CO2 corrosion therefore becomes a priority and it can become expensive. The replacement of the original Forties MOL and the severe damage to the Beatrice MOL are two examples of high costs that BPX have incurred in recent years due to unpredicted corrosion rates. Successful management of CO 2 corrosion starts off with the identification of risks and continues with the provision of suitable controls and the review of the success of the controls via monitoring - as illustrated in Figure 2.

INTRODUCTION

Figure 2: The Feedback Loop that is Required for Successful Management of CO2 Corrosion

Quantify Risk

Apply Controls

Monitor Effectiveness

This document sets out BPs approach to the quantification of CO2 corrosion risk through the use of predictive models. In doing so, it also discusses the reliance that can be placed on corrosion inhibition as the only viable control measure for carbon steel and the importance of suitable corrosion monitoring. To put the importance of this into context, corrosion costs BPX 8.3% of its capex budget and increases lifting costs by 14%, an average of over 8 cents per barrel. Figure 3 shows that the costs are distributed across the entire range of facilities.

INTRODUCTION

Figure 3: The Distribution of Costs of Corrosion Across Ten BPX North Sea Assets, 1990 to 1994.

Topsides 23%

Personnel 1%

Downhole 13%

Chemicals 4%

Subsea 59%

The quantification of corrosion risk is required at several stages during an assets life. The most obvious period is during the project phase when the original materials of construction are being selected. This process must be repeated during the life of the asset if failures or expansions require the procurement of additional facilities. Quantifying the corrosion risk is also important in tailoring inspection strategies. Risk based inspection is now widely adopted and, as CO2 corrosion represents one of the most important factors governing the probability of failure for much equipment, a reasoned approach should be taken. It is important that this approach is theoretically sound but also reflects past experience. This version of the BP CO2 prediction model is the first to be published since 1993/4 when the guidelines on multiphase and wet gas transport respectively were issued. The new guidelines incorporate changes by the authors to the semi-empirical model used in the original guideline as well as comprehensive guidance on how to use the spreadsheet included with this version. The new model also includes the ability to predict the affects of changing flow velocities on uninhibited corrosion rates.

INTRODUCTION

The new guidelines also consider the probabilistic approach to predicting CO2 corrosion. Probabilistic approach to design in general is becoming more widespread and offers several advantages over the traditional deterministic approach. The probabilistic approach is neither endorsed nor disallowed but is discussed as, in some cases, it may be more appropriate than a deterministic approach. The approach to designing for the use of corrosion inhibitors has been changed significantly. The previous approach described the affects of an inhibitor through the use of an efficiency factor, such as 90%. This does not reflect BPXs recent field data generated under severe conditions which showed inhibitors can be more effective than predicted. "Inhibitor efficiencies" have therefore been replaced with "inhibitor availabilities" that more closely reflect field experience. There is a general move in the industry towards this methodology and it offers several advantages. However, it has become clear that for inhibitors to work effectively the corrosion management system must be highly organised. Recommendations are therefore included on methods to ensure that the inhibitor availabilities assumed at the design stage occur during the operational stage.

"Cassandra 98" Corrosion Prediction Spreadsheet


by A J McMahon

INTRODUCTION "Cassandra 98 is BP's new implementation of the 1991, 1993 and 1995 CO2 corrosion prediction models published by De Waard et al. The pocket inside the front cover of this report contains a floppy disc with the programme together with a set of installation instructions. The 1991 and 1993 De Waard models are already widely used within BP and elsewhere in a variety of customised forms. This report describes the new Cassandra 98 spreadsheet for Microsoft Excel. It is based primarily on the 1993 De Waard model, incorporates some equations from the 1991 model, and uses the 1995 model to assess velocity effects. The spreadsheet is intended to capture all the best features of the 1991, 1993 and 1995 models [1,2,3]. Certain extra features from outside the De Waard papers, based on standard physical chemistry, have also been included. The source, background and limitations of all the assumptions and equations in the spreadsheet are fully documented in these guidance notes. The Cassandra 98 spreadsheet is written in a simple and accessible format within Microsoft Excel (version 7.0). It avoids the use of macros or special techniques so that the logic and the calculations are as transparent as possible. This approach also ensures that the spreadsheet is immediately compatible with new versions of Excel. The Excel add-in module "CRYSTAL BALL" (from Decisioneering Ltd, 1380 Lawrence Street, Suite 520, Denver, Colorado 80204, USA. Tel: +1 303 292 2291. Cost ~100) enables probability distributions to be set for each input cell and it then uses Monte-Carlo simulation to combine these into a probability distribution for the resulting corrosion rate. You must buy "CRYSTAL BALL" separately for your Excel environment. It can't be bundled with this spreadsheet. The detailed use of CRYSTAL BALL is well covered in the manufacturer's handbook and therefore is not repeated in these guidelines. Care is required when comparing the output of any existing in-house version of the De Waard models against this new Cassandra 98 spreadsheet. It is very easy for errors and untested assumptions to be entered into a spreadsheet which might then perhaps be passed on from user to user and often compounded with other assumptions. Cassandra 98 has been written from scratch with a detailed re-evaluation of all assumptions, all of which are presented. Cassandra 98 is intended to be a standard, reference version of the De Waard approach for use within BP and its partners, until such time that a more consistent approach to corrosion modelling becomes established within the oil industry. The activities of the NORSOK industry forum in Norway are making helpful moves in this direction. 5

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

QUICK START This section gives enough information to allow experienced modellers to make a start. The subsequent section gives a more detailed description of all the input and output parameters. The spreadsheets themselves also carry frequent "cell notes". These are marked by a red dot in the top right hand corner of those cells. Just double click on the cell to read the contents.
Input Parameters

To carry out a basic calculation enter the following input values into the cells with a white background:
Parameter P %CO2 %H2S water composition brine pH Comments Units Cell F7 F8 M8 A15-L15 F17

Table 1: Input parameters for a numeric calculation

T Ts

d U

total gas pressure bar CO2 in gas mole % (NB = v/v%) H2S in gas mole % (NB = v/v%) ion ppm values ppm (NB = mg/ltr) enter known value, or enter "d", "o", or "x" to accept one of the calculated values shown in F18-F20 (see Page 17) oC System temperature oC Scaling temperature, enter the calculated scaling temperature, given in cell F26, or another known or preferred value hydraulic diameter m velocity m/s

F24 F25

M24 M25

Probabilistsic Inputs

Only the inputs in the preceding Table are needed for a straightforward numeric calculation. Some further information is required in order to carry out a probabilistic calculation using CRYSTAL BALL. The spreadsheet can easily be customised by individual users to permit more extensive handling of probabilities:

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

Table 2: Additional Input Parameters for a Probabilistic Calculation

Parameter P %CO2 brine pH

Cell F7 F8 F17

Comment use a uniform distribution; set F7 as the maximum; set G7 as the minimum use a normal distribution; adjust standard deviation as necessary must enter a known or a calculated value; use a normal distribution; adjust standard deviation as necessary use a uniform distribution; set F24 as the maximum; set G24 as the minimum use a uniform distribution; set M24 as the maximum; set N24 as the minimum use a uniform distribution; set M25 as the maximum; set N25 as the minimum

T d U

F24 M24 M25

Output Parameters

The resulting output parameters are described in Table 3. See p23 for a more detailed description of how to interpret and use these values. Briefly, the 1993 rate should be regarded as the minimum. Velocity effects may increase this minimum rate as shown by the 1995 rate. Hence, the 1993 and 1995 rates will normally give the lower and upper bounds on the expected corrosion rate. The 1995 model is not accurate at low velocities and so it should be ignored whenever it falls below the 1993 value.
Parameter 1993 basic Vcor 1993 correction factors 1993 corrosion rate 1995 corrosion rate Cell E32 Comments

Table 3: Output Parameters

93/95 merged corrosion rate

the uncorrected corrosion rate for static conditions G32-K32 correction factors for pH, fugacity, scaling, and glycol G34 the corrosion rate for static conditions corrected for pH G39 the corrosion rate for dynamic conditions calculated from the components Vr and Vm in G37 and G38 G41 the average of the 93 and 95 corrosion rates; this cell enables "CRYSTAL BALL" to combine the 93 and 95 probability distributions

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

LIMITATIONS OF CORROSION PREDICTION MODELS

The use of simple equations and the precision of the spreadsheet environment can lead one to think that the De Waard corrosion models are equally precise. However, this is not the case. The models are only valid over a certain range of conditions, and even within this range a certain amount of data has been ignored if it doesn't fit the main trends. Each model appears to be constructed by obtaining a large number of corrosion rates over a range of conditions and then finding an equation which draws a line passing close to the majority of this cloud of points. The equations appear to be freely adjusted in order to give the best fit to the data. The primary concern is to obtain a good fit to the data, rather than obtaining mechanistically rigorous equations. These are empirical engineering models rather than scientific theories. Neither the 1991 or 1993 De Waard papers give many precise details about the range of validity of the models. The 1995 paper does give a more thorough set of figures (see below) but still omits important features such as the type of brine used in the tests, and the elapsed time when the corrosion rates were measured. De Waard's very early work used a 0.1% NaCl solution [4] and this may well have been used in all the subsequent studies because his main focus has always been low salinity water in gas lines. Table 4 shows the approximate ranges of validity for the different parameters in the Cassandra 98 spreadsheet.
Table 4 : Range of Validity of De Waar d Models

Parameter

Range of 1991 Range of 1995 & 1993 Model Models <200 bar <10 bar --<140oC 0 m/s not defined 0.3-6.5 bar --20-80oC 1.5 -13 m/s

Comments

P fCO2 Oddo & Tomson pH XLpH T U

<200oC, <1000 bar <120oC

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

The spreadsheet gives freedom to enter any value for most parameters. When the input value is outside the approximate range of the 1991 and 1993 De Waard models then the text will turn RED in the cell as a warning. The predicted corrosion rate may still be useful but the user must accept the additional risk of going beyond the known limits of the correlations.
Limits of the 1995 Model

To develop the 1995 model [3] corrosion rates were obtained on the IFE flow loop (Kjeller, Norway) using a radiochemical technique to measure corrosion rates. Tests were carried out over 2-3 days but there is no information about the corrosion rate profile over this time or when the final data point was taken. Data were obtained for the following conditions. St-52 DIN 17100 steel (Cr 0.08%, C 0.18%) which is similar to ASTM A537 Gr1 0.1, 3.1, 8.5, 13 m/s flow velocity 20 - 90 oC 0.3 - 20 bara CO2

Certain inconsistencies in the data set were eliminated prior to developing the model. These included: 0.1 m/s excluded 13 m/s excluded when corrosion rate less than at 8.5 m/s 90oC excluded CO2 >6.5 bar excluded

Eventually 221 data points were used in the main correlation (Figure 2 ref 3). The main equations are specific to St-52 steel because, "The equations obtained for St-52 showed a complete lack of correlation for the other steels". The 15 other steels were normalised steels and quench-and-tempered (Q&T) low alloy steels. These were examined over the following conditions to produce some modified equations which take account of steel composition. 3.1, 8.5, 13 m/s flow velocity 60oC ca 2 bar CO2 pH 4,5,6

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

For normalised steels a "Cr correction" and a "C correction" can be calculated separately and together. For Q&T steels the "C correction" has no effect and only the "Cr correction" is relevant. The Cassandra 98 spreadsheet does not include the steel composition equations due to the poor correlations obtained when fitted to the model.
Errors on Corrosion Rates

Errors in matching equations to data points are defined in the 1995 paper by "coefficients of determination". This is a complicated statistical function ranging from 0 (poor correlation) to 1 (perfect correlation). It is not the same as the "correlation co-efficient" in regression analysis which scales from -1 to 1. The "co-efficients of determination" in the paper are 0.91 for the main St-52 equations (after excluding the data that doesn't fit), 0.83 for the normalised steels, and 0.80 for the Q&T steels. For the main St-52 correlation this corresponds to a standard deviation of 25% on the predicted corrosion rate. This is the error given in this spreadsheet. Because of this error the predicted corrosion rates are only shown to one decimal place. A "CRYSTAL BALL" probabilistic analysis gives a more realistic impression of the error on each prediction.

APPLYING THE MODEL TO DIFFERENT FIELD SITUATIONS The De Waard models were all developed using water-only systems in the laboratory. The 1993 model is intended for nearly static, aqueous conditions and so for all but the lowest velocities (see page 77) it can be regarded as the minimum corrosion rate of a water wet region in a gas/water, water/oil, or a water/oil/gas system. Due to the different hydrodynamics in these field cases some assumptions are required in order to apply the 1995 model effectively. These assumptions will only affect the diameter and velocity values used as inputs in the model. The other inputs will be unaffected. Table 5 gives some suggested assumptions. However, users are free to develop their own approaches to meet the demands of their own particular circumstances. Some of the issues involved in extrapolating the models to the field are discussed in more detail on pages 27-28.

10

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

Table 5: Applying the 1995 De Waard Model to Field Situation

Field Situation Water only Liquid/Gas

Recommended Approach use pipe diameter and water velocity use hydraulic diameter (see p 21) use true liquid velocity rather than nominal velocity (see p 22) use pipe diameter and total liquid velocity (n.b. this ignores the possibility of water drop out or stratification which could lead to the water phase moving more slowly than the oil phase) use a specialist multiphase program to calculate the wall shear stress or the "C factor" for the pipe system, then choose diameter and velocity inputs which reproduce this hydrodynamic value.

Water/Oil

Water/Oil/Gas

DETAILED DESCRIPTION OF THE SPREADSHEET

Units are specified for each parameter listed in this section. The same units are assumed in all the equations given below and throughout the Cassandra 98 spreadsheet. The spreadsheet has a "units conversion box" at cell P5. The UNITS spreadsheet allows conversions between a wider range of units. The SALTS spreadsheet enables conversion between an ionic analysis of brine and the salts required to prepare a synthetic analogue. The FUGACITY spreadsheet is a data-base used to calculate fugacity corrections at high total pressures.
Total Pressure

P...total gas pressure (bara, i.e. bar absolute)

INPUT

cells F7 and G7

For a multiphase system this is simply the prevailing local P in the gas. For a liquid only system it is the P in the last gas phase which was in equilibrium with the liquid, e.g. the separator gas in the case of a crude oil export line. For a downhole liquid pressurised above the bubble point then use the bubble point pressure (Figure 4). For a simple numeric calculation, enter the P value into cell F7. Cell G7 is then unused. For a probabilistic calculation using "CRYSTAL BALL", set up a uniform distribution for P with F7 set as the maximum and G7 as the minimum. 11

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

Figure 4: Schematic Diagram of an Oil Production System (downhole, separator, export)

%CO2

%CO 2 ...CO2 in gas (mole%, which is same as v/v%)

INPUT

cell F8

For a multiphase system this is simply the prevailing local %CO2 in the gas. For a liquid only system it is the %CO2 in the last gas phase which was in equilibrium with the liquid, e.g. the separator gas in the case of a crude oil export line. For a downhole liquid use the %CO2 in the gas formed at the bubble point. If this gas analysis is not available then use the CO2 dissolved in the brine, the Henry's constant, and the bubble point pressure to backcalculate the "effective %CO2" which would be required in the bubble point gas in order to sustain the known level of dissolved CO2 (see box at cell P19). Indeed, this procedure can be followed for any region where the CO2 dissolved in the brine is known, but the gas analysis is unknown. There may be occasions when it is helpful to apply parts of the Cassandra model to a water which is in equilibrium with ambient air (e.g. for pH predictions). The appropriate atmospheric inputs are P = 1 bara and %CO2=0.035 mole%. Remember that under these conditions the corrosion prediction from the model will only relate to the dissolved CO2 component and not the dissolved O2. For a probabilistic calculation using "CRYSTAL BALL", set up a normal distribution for %CO 2 using an appropriate standard deviation.

12

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

pCO2

pCO 2 ...partial pressure of CO2 (bara) pCO2 = P.%CO 2 100 OUTPUT

OUTPUT

cell F9

fCO2

fCO 2 ...fugacity of CO 2 (bar)

cell F10

The non-ideality of gases means that at high total pressures the partial pressure is not an accurate description of the activity of a gas component. The fugacity is the true activity of the gas component. The 1991 and 1993 models use pCO 2 in the main corrosion prediction equations and then at the end apply a fugacity correction factor (Ffug) to account for fugacity effects. In Cassandra 98 the equations from the 1991 and 1993 models use fCO2 directly, therefore there is no need to use a fugacity correction factor (Ffug). The equations from the 1995 model in Cassandra 98 also use fCO2 directly instead of pCO2. Hence, in Cassandra 98, it is fCO2 which is used as the primary parameter for all the equations which consider CO2 as an input. Fugacity data from the work of R H Newton [5] are tabulated in the FUGACITY.XLS spreadsheet in the workbook. The Cassandra 98 spreadsheet uses the input values of temperature and total pressure to look-up the correct value of the fugacity co-efficient () in the FUGACITY spreadsheet, fCO2 = pCO 2 The R H Newton data are generally applicable to many pure gases. The data show fugacity co-efficients as a function of "reduced temperature" and "reduced pressure", Tr = T Tc

where

Tr is reduced temperature (dimensionless) T is the prevailing local temperature (oC) Tc is the critical temperature for the gas (from tables) (oC)

13

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

Pr =

P Pc

where

Pr is reduced pressure (dimensionless) P is the total pressure (bar) Pc is the critical pressure for the gas (bar)

Oilfields produce gas mixtures rather than pure gases. Hence, a difficulty arises in deciding whether it is the Tc and Pc for methane or for CO2 that one should use. In the Cassandra 98 spreadsheet, empirical values of Tc and Pc are assumed which allow the Newton model to agree with the CO2/methane mixed gas fugacity data in Figure 5 of the 1993 De Waard paper to 10%. In other words the De Waard data are used to calibrate the Newton model.
Table 6: Reduced Temperature and Reduced Pressure Values for CO2 and Methane

Tc (oC) CO2 methane empirical values used to correlate with De Waard data 31 -82 -37

Pc (bar) 73 45.8 56.7

The De Waard calibration data are valid up to 140oC and 250 bar. The Newton data extends beyond these levels up to 300oC and 400 bar. The general trends in the data will be accurate under these extreme conditions, however, the absolute values are unchecked. For accurate work it will be necessary to calculate or obtain the correct value of fugacity from elsewhere and then manipulate %CO2 in cell F8 by trial and error in order to obtain the correct fugacity in cell F10.
%H2S

%H 2 S...H2S in gas (mole%, which is same as v/v%)

INPUT

cell M8

H2S is not included in any of the De Waard models. It is only used in the Cassandra 98 spreadsheet in the calculation of solution pH by XLpH (see below). It can be ignored completely simply by entering zero.

14

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

It is by lowering the solution pH that H2S can potentially increase the corrosion rate, often in synergy with CO2. In practise, H2S tends to promote FeS surface films which reduce the observed general corrosion rate but which increase the likelihood of localised corrosion whenever the film fails. The CO2 general corrosion rate is often assumed as the worst-case localised corrosion rate for the regions with no FeS film. An alternative approximate approach for handling the presence of H2S is to assume that every 1 mole% H2S has the same corrosivity as 0.01 mole% CO2. This rule of thumb assumes that 1 ppm dissolved CO2 and 200 ppm dissolved H2S give roughly equal corrosion rates [6], and that H2S is roughly twice as soluble in water as CO2 for a given partial pressure [7].
pH2S

pH2S ...partial pressure of H2S (bar) pH2S = P . %H2S

OUTPUT

cell M9

LIQUID PARAMETERS
Water Chemistry

water chemistry ..ion concentrations (ppm, same as mg/ltr) INPUT cells A15-L15 The water chemistry is used to calculate the solution pH (see below). Enter ppm values for Na+, K+, Ca2+, Mg2+, Ba2+, Sr2+, Cl-, HCO3-, SO42-, Fe2+, acetate. (NB enter the sum of all organic acids as acetate). Enter the %v/v value for glycol in cell L15. Use the SALTS spreadsheet to check that the total positive and negative charges of the ions are roughly balanced. Any significant misbalance (e.g. >10%) may invalidate the pH calculation. Note that ion charges are handled in general chemistry by using the term "equivalents": 1 mole of positive charges is equal to one equivalent; in other words 0.7 mole of Ca2+ ions is equal to 1.4 equivalents of positive charge. Some further aspects of the acetate entry are discussed on p.19.

Total Dissolved Solids

TDS...total dissolved solids in water phase (ppm, same as mg/ltr) OUTPUT cell M17

15

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

This is the sum of all the individual dissolved ions concentrations. TDS and [HCO3-] are used in the Oddo & Tomson pH calculation. TDS is also used to estimate the "salting-out" of CO2 as salinity increases. This will tend to reduce the concentration of dissolved CO2 and thereby reduce the corrosion rate [8]. The box at X19 shows how to apply the salting-out correction. The procedure uses "Henry's Law" to calculate the solubility of a gas in a liquid. pCO2 = K H XCO2 where KH is Henry's constant (bar/mole fraction) XCO2 is mole fraction of CO2 dissolved in brine.

The Henry's constant from the De Waard paper is only valid for a low salinity brine (ca 0.1% NaCl). Therefore, by calculating the true Henry's constant for a specific brine it is possible to apply a salinity correction to the De Waard corrosion rate. The salt-correction procedure first calculates the Henry's constant used by the De Waard model (equation 28 from the 1993 paper- which is used in the derivation of equation 13 in the 1993 paper), log10 K H = 1088.76 5.113 T + 273

where KH is Henry's constant (mole/ltr bar) Note that this KH equation from the De Waard paper has different units (mole/ltr bar) from those given earlier (bar/mole fraction). Much of the confusion over Henry's constants arises from the wide and sometimes awkward range of units which can be used to express the parameter. For consistency in this report the De Waard equation for an aqueous solution can be rewritten in order to maintain KH in units of (bar/mole fraction).. 1088.76 18 log10 K H = 5.113 T + 273 1000

where K H is Henry's constant (bar/mole fraction)

16

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

The true Henry's constant is a function of both salinity and temperature (Appendix 1) so that, K true (for 0 125C) = (1.77 T + 47.1) H TDS + (45.2 T + 559) 10000

K true (for 125 200C) = 250 H

TDS + 6500 10000

Therefore, the salt-correction factor, Fsalt, is, Fsalt = KH


De Waard

K true H

The best way to use Fsalt is to apply it to fCO2 to give an "effective CO2 fugacity". This "effective fCO2" will give the correct dissolved CO2 concentration when used with the other equations in the Cassandra 98 model. The salt correction effect only becomes significant for TDS > 10% w/v.
Brine pH

pH...brine pH control parameter

INPUT

cell F17

Enter the known pH value, or else enter a letter to accept one of the calculated pH values given in cells F18, F19, or F20

r "d" or "D" will accept the De Waard distilled water pH r "o" or "O" will accept the Oddo & Tomson brine pH r "x" or "X" will accept the BP XLpH calculated value.
The accepted value is displayed in cell F21 for confirmation.

17

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

When doing a probabilistic calculation using CRYSTAL BALL then a numeric value of pH (either known or calculated) must be entered. Use a normal distribution for the probability adjusting the standard deviation so as to cover appropriate minima and maxima.
pH(CO2)

pH(CO 2 )...pH of distilled water containing CO2 Equation (8) from the 1995 paper...

OUTPUT

cell F18

pH(CO2) = 3.82 + 0.000384 T - 0.5 log10 (fCO2) fCO2 is used here rather than the pCO2 quoted in the original paper. The equation is valid over 10-80oC. It gives the pH for pure water containing dissolved CO2 at the prevailing temperature and fCO2.
pH(act)

pH(act, Oddo) ..Oddo & Tomson calculated pH in brine OUTPUT cell F19 An empirical equation from reference 9...

HCO 3 + 8.68 + 0.00405 (T * 9 / 5 * 32)... pH = log10 fCO2 * 14.5 * 61000 +0.000000458 (T * 9/5 * 32)2 - 0.0000307 (P * 14.5)... TDS 0.477 58500
1 /2

TDS + 0.193 58500

fCO2 is used here rather than the pCO2 quoted in the original paper. The equation is valid up to 200 oC and 1200 bar, but is inaccurate for low values of [HCO3-]. The Cassandra 98 spreadsheet is set to give an error for pH(act, Oddo) if [HCO 3-] < 50 ppm.

18

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

pH(act, XLpH)

pH(act, XLpH) ...XLpH calculated pH in brine

OUTPUT

cell F20

XLpH is an Excel add-in function for calculating both pure water and brine pHs with no restrictions on salinities or component concentrations. It was developed by XTP, Sunbury using well documented code published by the US Geological Survey (the "PHREEQ" model). The original version of XLpH [10] has since been updated to include pH2S as an input parameter. XLpH has been validated against other pH models such as in CORMED and also against literature and recent laboratory values. XLpH uses the individual ion concentrations in cells A15-L15. The positive and negative charges must be approximately balanced (see "water chemistry", p15, above). XLpH will automatically compensate for any small misbalances by adding Na + or Cl - ions. Enter the sum of all organic acids as acetate. Note that the pH of CO2containing-brine will differ depending on whether the acetate is added in the form of sodium acetate salt or acetic acid... pH of 0.5 M NaCl / 300 ppm NaHCO3, 1 bar CO2, 25 oC plus... no acetate 5.53 6.8 mM Na acetate (i.e. 571 ppm) 5.41 6.8 mM acetic acid (i.e. 422 ppm) 4.17

XLpH assumes that the acetate value entered in cell K15 is acetic acid, because this is the worst case. If one wishes to assume Na acetate then zero should be entered for Ac and the molar equivalent of Na acetate should be added to the Na and Cl entries. Unfortunately a field water analysis will not directly reveal whether Na acetate or acetic acid should be used to simulate the water chemistry. This can only be established by making laboratory pH measurements under CO2 saturation and comparing the results with the XLpH model. Inclusion of the organic acid concentration will always improve the reliability of a prediction. However, when organic acid data is not available it is possible to make some rule-of-thumb approximations in order to aid progress. Organic acids are typically present in formation water at <30ppm. Therefore, for bicarbonate >150ppm, the presence of organic acids is likely to make little difference to the calculated pH and therefore corrosion rate. In such cases, an API water analysis (which omits organic acids) will often suffice. If the formation water is low in bicarbonate (<150ppm), then there is more chance that organic acids could make a significant contribution to the in situ pH and calculated corrosion rate and so an acetate entry should be added to the water analysis. 19

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

Accepted pH

accepted pH ...confirmation of selected pH

OUTPUT

cell F21

This is confirmation of the pH value which has been accepted for the corrosion prediction equations.
T

T...temperature (oC)

INPUT

cell F24

The prevailing local temperature. When doing a probabilistic calculation using CRYSTAL BALL then use a uniform distribution for the temperature : set F24 as the maximum and G24 as the minimum.
Scaling T

T s ...selected scaling T (oC)

INPUT

cell F25

Enter a preferred value for the scaling temperature or enter "a" (or "A") to accept the calculated value shown in cell F26. Researchers are still actively investigating the issue of what happens to corrosion rates at temperatures above the scaling temperature. Previous work has shown that sometimes the scale films are protective and can reduce the corrosion rate, whereas sometimes the films are non-protective so that the corrosion rate continues to increase. Choosing one or other of these options could on the one hand lead to significant under-design, and on the other hand to significant over-design. Therefore, until the matter is fully resolved BP prefers to choose a middle course for design purposes. BP assumes that the corrosion rate reaches a peak at the scaling temperature and remains on a plateau at the same value for higher temperatures. The Cassandra 98 spreadsheet follows this approach. In order to achieve this outcome both fCO 2 and pH are set to a plateau for T > Ts.
Figure 5: The Possible Effects of High Temperature Scaling on the Corrosion Rate

Corrosion Rate

IFE, Norway data BP approach

De Waard approach Ts

Temperature

20

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

De Waard Calculated Scaling Temperature

Ts ...De Waard calculated scaling T ( oC) Equation (13) from the 1995 paper,

OUTPUT

cell F26

2400 Ts = 273 6.7 0.44log10fCO2

This is obtained by setting log 10 Fscale = 0 (i.e. F scale = 1) in equation (13) in the 1995 paper. Note that the equation above is expressed in oC and uses fCO2 rather than the oF and pCO2 used in the paper. The 1993 paper gives a similar equation to the 1995 paper but uses a factor of 0.67 in front of the log term instead of 0.44.
Diameter

d...hydraulic diameter (m)

INPUT

cell M24

A diameter input value is only required for the velocity equations in the 1995 model. It is not needed for the 1993 model. The 1995 paper actually uses "hydraulic diameter" rather than a simple pipeline diameter. Let Dp be pipeline diameter, and let Dh be hydraulic diameter, then, ..for gas/liquid pipelines, Dh < D p Dh = 4 A / S ..where A is the cross-sectional area of the liquid in the pipe S is the cross-sectional perimeter length of the liquid region (i.e. liquid/pipe + liquid/gas interfaces, see Figure 6)

..therefore for a pipeline full of liquid, Dh = D p

21

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

Figure 6: Explanation of Parameter "S" in a Gas/Liquid System

cross-sectional perimeter length of the liquid region

There is a box at cell P39 for calculating hydraulic diameters in gas/liquid lines. The ratio of the liquid and gas cross-sectional areas, or the ratio of the liquid depth to the pipe radius, is required as an input parameter. Calculation of this parameter is outside the scope of the Cassandra 98 spreadsheet. When doing a probabilistic calculation using CRYSTAL BALL then use a uniform distribution for the hydraulic diameter : set M24 as the maximum and N24 as the minimum.
Flow Velocity

U...flow velocity (m/s)

INPUT

cell M25

A flow velocity input value is only required for the velocity equations in the 1995 model. It is not needed for the 1993 model. There is a box at cell P5 which enables calculation of flow velocity from pipe diameter and flow in liquid only lines. The calculation is more complicated for the liquid phase in gas/liquid lines, therefore, the box at cell P39 should be used. When doing a probabilistic calculation using CRYSTAL BALL then use a uniform distribution for the flow velocity : set M25 as the maximum and N25 as the minimum.

22

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

Outputs : 1993 De Waard Model


Vcor...Basic Corrosion Rate

Vcor ...basic corrosion rate (mm/yr.) Equation (13) from the 1993 paper,

OUTPUT

cell E32

log10 Vcor = 7.96

1710 0.67 log10 (fCO2 ) T

The basic corrosion rate is adjusted by multiplying with the pH and occasionally the glycol correction factors (FpH and Fglyc respectively). The application of each of these is discussed below. For the basic corrosion rate and the correction factors, the values reached at the scaling temperature are set to remain the same at higher temperatures. This is to ensure that the corrosion rate reaches a peak at the scaling temperature and then remains on a plateau at the same value for higher temperatures (see Ts section above). Hence, the BP approach does take account of scaling at high temperatures but doesn't use the De Waard scaling factor, Fscale, directly.
pH Correction Factor

F pH ...pH correction factor

OUTPUT

cell G32

Equations (9) and (10) from the 1991 paper, log10 FpH = 0.32 (pH CO2 - pH act) for pH CO2 > pHact where ...pHact is the actual pH of the brine which wets the pipewall ...pHCO2 is the pH under the same conditions but in pure, salt-free water

log10 FpH = - 0.13 (pHact - pHCO2)1.6 for pHCO2 < pHact

23

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

These equations show that as pHact rises, FpH will get smaller and so the corrosion rate will fall. These equations use pHCO2 instead of the "pHsat" used in the De Waard paper. pH sat is the pH at which a brine first becomes saturated with either FeCO3 or Fe3O4 as a result of the steel corroding and building up dissolved Fe2+ in the solution. The problem with pHsat is that it is difficult to define. Even the De Waard paper only gives some approximate expressions for one particular brine composition (10% NaCl). Furthermore, there is serious doubt over the whole concept of a fixed saturation pH due to the observation of massive supersaturation effects by IFE (Norway) and also within BP. Dissolved Fe2+ concentrations can often reach hundreds of ppm and can exceed the theoretical saturation values by orders of magnitude. Hence, pHsat is not a reliable concept. Until the pHsat issue is resolved BP prefer to use pHCO2 as an alternative reference point. It has the advantage that it is well defined and is valid over a wide range of conditions. Therefore, a pure water system will give pHact = pH CO2 and so FpH = 1 in the BP approach. Certain conditions can make pHact < pH CO2 (e.g. high salinity, zero bicarbonate) and so FpH > 1. The presence of bicarbonate will tend to make pHact > pHCO2 and so F pH < 1. One way of reconciling these divergent approaches is to say that the direct De Waard approach uses Fph to derive the initial corrosion rate in a brine before corrosion products build up and gradually reduce the corrosion rate until it reaches a steady state. This is the issue discussed in the 1993 De Waard paper. The BP approach on the other hand does not deal with initial corrosion rates at all. It deals only with steady state corrosion rates and uses Fph to express the effect of water composition on the steady state rate. This effect is not covered in the direct De Waard approach. In essence BP have taken an equation from the direct De Waard approach and then adapted it for another purpose. Hence, overall, the two approaches are different but consistent.
Fugacity Correction Factor

F fug ...fugacity correction factor Equation (3) from the 1991 paper,

OUTPUT

cell J32

1.4 log10 Ffug = 0.67 0.0031 P T + 273

24

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

Ffug is not required in the BP approach because fCO2 is used in preference to pCO2 throughout the calculation and so fugacity has already been accounted for.
Scaling Correction Factor

F scale ...scaling correction factor Equation (16) from the 1993 paper,

OUTPUT

cell K32

1 1 log10 Fscale = 2400 T + 273 Tscale + 273

where

... T > Ts otherwise F scale = 1 ... Tscale is scaling temperature (defined above)

This factor is not used directly in the BP approach. It is included in the spreadsheet only for completeness.
Glycol Correction Factor

F glyc ...glycol correction factor

OUTPUT

CELL H32

Equation (20) from the 1993 paper, log10 Fglyc = A (log 10 W - 2) where ... A is a constant = 1.6 to a first approximation ... W is water content (%) of water/glycol mixture

BP only use this factor for cases without corrosion inhibitor. When a corrosion inhibitor chemical is used or is planned then BP assume that any effect of glycol is included within the corrosion inhibitor efficiency (normally 90%, but see discussion on pages 42-48).
Corrected Corrosion Rate

V' cor ...corrected corrosion rate (mm/yr.)

OUTPUT

cell G34

This is BP's preferred output from the 1993 DeWaard model. It is the base corrosion rate multiplied by the FpH correction factor. Note that for the basic corrosion rate and the correction factor, the values reached at the scaling temperature are set to remain the same at higher temperatures. This is to ensure that the corrosion rate reaches a peak at the scaling temperature and then remains on a plateau at the same value for higher temperatures (see T(s) section above). Hence, the BP approach does take account of scaling effects at high temperatures but doesn't use the De Waard scaling factor, Fscale, directly. 25

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

Outputs : 1995 De Waard Model The 1995 De Waard model is derived in a different fashion from the 1993 model, in particular it does not use the idea of correction factors applied to a base corrosion rate. Instead, the overall corrosion rate is calculated from two components : the reaction rate Vr and the mass transfer rate Vm.
Reaction Rate

Vr ...reaction rate (mm/yr.)

OUTPUT

cell G37

Equation (11) from the 1995 paper, log10 Vr = 6.23 1119 + 0.0013 T + 0.41log10 (fCO2 ) 0.34pH act T + 273

Mass Transfer Rate

Vm ...mass transfer rate (mm/yr.)

OUTPUT

cell G38

Equation (10b) from the 1995 paper, U 0.8 fCO2 d 0.2

Vm = 2.45

Overall Corrosion Rate

Vcor ...corrosion rate (mm/yr.)

OUTPUT

cell G39

Equation (2) from the 1995 paper, 1 Vcor = 1 1 + Vr Vm

where

Vcor is overall corrosion rate Vr is reaction rate Vm is mass transfer rate

26

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

1993 & 1995 Merged Corrosion Rate

Vcor ...merged corrosion rate (mm/yr.)

OUTPUT

G41

The merged rate simply takes the average of the 1993 and 1995 values. This allows CRYSTAL BALL to combine the probability distributions for the 1993 and 1995 rates so that one can see the lower and upper bounds on the expected corrosion rate.
1993 1995 merged V cor + Vcor Vcor = 2

The 1993 rate is regarded as the minimum. Velocity effects may increase this minimum rate as given by the 1995 value. The 1995 model is not accurate at low velocities so it is ignored whenever it falls below the 1993 value, and then the merged rate is the same as the 1993 rate. COMPARING OUTPUT FROM THE Cassandra 98 MODEL WITH FIELD DATA The validity of any corrosion prediction model depends on how well it agrees with the measured corrosion rates in the field. However, the comparison is not always straightforward. This is because the models are developed from well characterised, clean and stable systems in the laboratory, and they are being applied to partially characterised, dirty, and variable systems in the field where the full operating history is not always known. This is no criticism of field activities. It is simply a fact of life of operations where the aim is to produce hydrocarbons, not to generate completely rigorous corrosion data. The discrepancies between the models and r eal field corrosion data which do exist arise because there are parameters in the field which the model can not take account of effectively, or at all, e.g. surface coatings (scales, corrosion products, biomass), crude oil wetting, local hydrodynamics, weld metallurgy. The industry generally regards the De Waard model as conservative compared to the field, i.e. it over-estimates the field corrosion rate. Much of this opinion is based on anecdotal and semi-quantitative evidence - often not published in the open literature - but it is confirmed by the occasional formal presentation [12]. BP is currently compiling a database of field corrosion data

27

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

from a variety of sources which will be used to assess the Cassandra 98 spreadsheet presented here. In the meantime Table 7 gives a comparison of the Cassandra 98 spreadsheet against new laboratory data; data which were not used in compiling the model. The final column shows whether the observed corrosion rate falls within 15% of the range encompassed by the 1993 and 1995 models and there is some agreement. However, the discrepancies show the pitfalls in trying to push the accuracy of the model too far. It is best used to gain order of magnitude estimates of corrosive situations rather than absolute corrosion rates to several decimal places.
Table 7: Comparison of Model Predictions with Laboratory Data

corrosion rate (mm/yr.) T U fCO2 observed 93 95 correct? (oC) (m/s) (bar) model model BP 1993 0.1% NaCl, 3 litre flow loop (15 mm ID) 25 1.9 1 5 1.1 5.8 yes 25 1.9 0.27 2.2 0.5 1.9 yes 35 1.9 0.27 3.4 0.7 2 no BP 1992 Forties brine, beaker test and 5 litre flow loop (15 mm ID) 50 0 0.88 2.5 1.5 0.1 no 50 1.2 0.88 2.5 1.5 3.2 yes CAPCIS Flow Project Forties brine, flow loop (10 mm ID) 25 3.2 1 1.8 0.6 3.3 yes 50 1.1 0.88 3.8 1.5 3.2 yes 50 1.7 0.88 4.1 1.5 3.9 yes 50 2.5 0.88 2.5 1.5 4.4 yes 50 3.2 0.88 4 1.5 4.7 yes CAPCIS Flow Project 3% NaCl, flow loop (10 mm ID) 25 3.2 1 6 1.2 7.7 yes 50 3.2 0.88 12.1 3.1 9.2 no 70 3.2 0.88 17.4 5.3 8.4 no 50 1.1 0.88 6.8 3.1 4.8 no 50 1.7 0.88 7.3 3.1 6.4 yes 50 2.5 0.88 8.6 3.1 8.1 yes

28

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

APPENDIX 1 : "Henry's Law" Constants for CO2 Dissolved in Brine "Henry's Law" describes the solubility of a gas in a liquid, pCO2 = K H XCO2 where KH is Henry's constant (bar/mole fraction) XCO2 is mole fraction of CO2 dissolved in liquid

Henry's constants are dependant on both temperature and salinity and they are easily found for CO2 dissolved in pure water [e.g. 13]. The data for brines is less extensive [14-16]. Figure 7 is compiled using data from all these sources. The reduced number of points at higher salinity are still sufficient to show that the data in the 0-10% region can be reliably extrapolated up to ca 30% NaCl. Note that the 16 and 31% data at 75 and 100oC are actually for MgCl2 in the original paper but have been plotted in Figure 7 at the equivalent ionic strength of NaCl.

Kh (bar/mol frac)

Figure 7: Henry's Law Constants as a Function of Salinity

14000 12000 10000 8000 6000 4000 2000 0 0 5 10 15 20 25 30 35 [NaCl] %w/w T (oC) 200 175 150 125 100 75 50 30 10

The lines in this figure can be represented by the following equations (to within 15%),

29

"CASSANDRA 98" CORROSION PREDICTION SPREADSHEET

K H (for 0 125C ) =(1.77 T + 47.1)

TDS + (45.2 T +559 ) 10000

K H (for 125 200C )250

TDS + 6500 10000

where K H is Henry's constant (bar/mole fraction) Cell AD31 in the spreadsheet uses these equations to calculate the true Henry's constant for the input values of T and TDS.

30

The Use of Corrosion PredictionModels During Design by D M E Paisley

Introduction The value and purpose of predictive corrosion rate models should be neither overlooked nor exaggerated. The models (of which CO2 models are one example) are tools for the Materials Engineer to use during materials selection studies. The models help to quantify the corrosion risk and to help assess the impact of various process or production scenarios. However, corrosion rate prediction models should always be used in conjunction with other tools such as life cycle costing as well as previous operational experience if the final materials selection is to offer the optimal balance between cost and reliability. As each project will have unique economic factors, materials selection should reflect these and the economic assessment will be as important as the corrosion modelling in the selection of the final materials. In-depth coverage of techniques such as life cycle costing and estimating values are beyond the scope of this document but both techniques are briefly covered in a previous publication [17]. Over the past few years, several design guidelines have been issued by BP for dealing with CO2 corrosion risks. Each document deals with a specific application. This more general document summarises all previous guidelines but can not deal with the specific issues to the level of detail possible in the individual guidelines. The previously issued guidelines are listed in Table 8.

31

THE USE OF PREDICTIVE MODELS DURING DESIGN

Table 8: Previously Issued Design Guidelines

Report Title

Authors

Report Number Issue Date ESR.93.ER.013 1/3/93

A corrosion philosophy for the I D Parker transport of wet oil and J Pattinson multiphase fluids containing A S Green. CO2 A corrosion philosophy for the transport of wet hydrocarbon gas containing CO2 I D Parker J Pattinson A S Green.

ESR.94.ER.016

28/8/94

Assessment of a top of line D Paisley versus bottom of line corrosion J Pattinson ratio for use in the design of S Webster wet natural gas pipelines The application of pH moderation as a means of corrosion control for wet gas pipelines D Paisley

Branch Report No 124 421

5/10/92

ESR.95.ER.042

10/4/95

The effects of low levels of D Paisley hydrogen sulphide on carbon R Gourdin dioxide corrosion: A review of industry practice and a guide to predicting corrosion rates

ESR.95.ER.073

22/6/95

A corrosion philosophy for the transport of wet oil and multiphase fluids containing CO 2 This was the first undertaking in recent years to document a BP approach to defining internal corrosion risks and the basic approach is still followed. It recommended the use of the de Waard and Milliams model to predict insitu corrosion rates along with a 90% corrosion inhibitor efficiency. Much of the work is still valid but it is in the areas of high temperature scaling, corrosion inhibitor efficiencies and impact of various flow regimes that the

32

THE USE OF PREDICTIVE MODELS DURING DESIGN

new guidelines differ. Most of the recommendations made in these guidelines have been reproduced or superseded in the present document and therefore the original guidelines are redundant. A corrosion philosophy for the transport of wet hydrocarbon gas containing CO2 This was a companion document to the guidelines on wet oil and multiphase systems. The basic approach was similar but this document dealt with the specific wet gas application. Most of the recommendations made in these guidelines have been reproduced or superseded in the present document and therefore the original guidelines are redundant. Assessment of a top of line versus bottom of line corrosion ratio for use in the design of wet natural gas pipelines Wet natural gas pipelines operating under stratified flow have two distinct corrosion environments : (a) the bottom of line which is continually wetted by condensed water, hydrate inhibitor and hydrocarbons, and (b) the top of line which is wetted intermittently by condensing liquids. The corrosion rate at the top of the line is lower than that at the bottom due to the more limited exposure to corrosive species. Predicting this rate is done by predicting the bottom of line rate using models in the normal way and applying a moderating factor for the top of line rate. Up to 1992, BP used a factor of 0.3, i.e. the top of line corrosion rate was predicted to be 30% of the bottom of line rate. When inhibitors are used to control the bottom of line rate, the top of line corrosion rate becomes the limiting rate as inhibitors are assumed not to protect against condensing corrosion. This report reviewed the top of line factor and recommended the adoption of a moderating factor of 0.1. For inhibitor efficiencies up to 90%, the top of line corrosion rate is therefore not the limiting rate. This approach is no longer valid since BP have moved away from the direct use of inhibitor efficiencies, as described later in this report. However, the assumption that top of line rates are 1/10th of the predicted uninhibited bottom of line rates can still be used. For applications were the 'top of line' corrosion rate is the faster rate (using the 0.1 moderating factor) then a more detailed evaluation should be carried out. Such a scenario does not lend itself to the use of simplified guidelines.

33

THE USE OF PREDICTIVE MODELS DURING DESIGN

The application of pH moderation as a means of corrosion control for wet gas pipelines This technique is not widely applicable but may find niche applications in highly corrosive wet gas lines utilising recycled glycol for hydrate control. It is covered in more detail on p75 but if this technique is of interest the full guideline document should be reviewed. The effects of low levels of hydrogen sulphide on carbon dioxide corrosion: A review of industry practice and a guide to predicting corrosion rates This document summarised how low levels of H2S influence corrosion rates dominated by CO2. The conclusion was that H2S at levels below the NACE criteria for sulphide stress corrosion cracking (ref MR0175, NACE Publications) reduces general metal loss rates but can promote pitting. The pitting proceeds at a rate determined by the CO2 partial pressure and therefore CO2-based models are still applicable at low levels of H2S. Where the H 2S concentration is greater or equal to the CO2 value, or greater than 1 mole%, the corrosion mechanism may not be controlled by the CO2 and therefore CO2 based models may not be appropriate. Summary of Previous Guidelines In summary, the old guidelines are generally still applicable. What has changed is BPs views on the reliability and performance of corrosion inhibitors as well as the availability of updated models incorporating flow affects. The old guidelines defined a corrosion inhibitor efficiency of 90% with no scope for variation. There were also stringent velocity restrictions for use under multiphase conditions which restricted the energy of slug flow to below 20 Pa, later raised to 100 Pa. In light of favourable field data, this approach is now seen as too pedantic and inhibitor availabilities are seen as a better way of describing the role of inhibitors. These differences in approach are covered in more detail in the following sections. Furthermore, the corrosion rate prediction model (p5-30) does not cover some aspects that are important during design and these are covered in the next section.

34

THE USE OF PREDICTIVE MODELS DURING DESIGN

Important Factors not Covered by the Corrosion Model


The Probabilistic Approach to Predictive Modelling

The modelling approach outlined in this document deals with all the inputs (mole% CO2, temperature etc.) on a deterministic basis. However, each input will have a level of uncertainty associated with it and this can have important effects on the outcome. One way to deal with this it to calculate a range of output values, (in this case the predicted corrosion rate) across the whole range of input values. Where the model is dealing with several inputs (temperature, pressure, CO2 mole %, pH, scaling factor), this can be time consuming. Also, the value of these inputs will not all vary in a uniform manner. Some will behave uniformly while others may behave in a normal or log-normal manner. Calculating the impact of all these variables is time consuming, unless a programme such as Crystal Ball is used. This is an add-in to Excel and handles the variability by performing a Monte Carlo analysis. Any number of iterations can be performed and the output is displayed in terms of a probability, rather than as a discreet value. In general, a minimum of 1,000 iterations, involving tens of thousands of individual calculations are required to show the effects of the variability in input data. A modern PC can perform such a task in a minute or two. The important factors to consider are the range and type of distribution assumed for each variable. If process data are available, this will form an ideal basis for determining the range and type of distribution but if this is lacking, some assumptions will have to be made. Using distributions to define variables in a predictive model can have significant effects on the outcome.

Worst Case Design

Engineering design traditionally uses worst case inputs so that the final design will be safe under all foreseeable combinations of events. This approach has also been adopted when predicting corrosion rates, where pressure and temperature etc. are used as inputs to the models. In the past this approach was the only viable one as predicting the enormous range of possible outcomes for all variables would have been too time consuming but it can result in substantial over-design. Metal loss corrosion processes do not lead to sudden failure due to a combination of variables over short time periods

35

THE USE OF PREDICTIVE MODELS DURING DESIGN

(unlike high pressure which can lead to an instantaneous failure) but rather reflect a combination of varying conditions over a longer time period. Using the worst case values is therefore not a sensible approach, if a range of more realistic values can be handled. In defining a range of likely operating variables such as temperatures and pressures, the design values will form the maximum for the respective distributions but lower values should be included. Defining this range will require inputs from the Process and Reservoir Engineers. Due to the nature of the uncertainty, such that all values within the range are as likely as each other, Uniform distributions are probably the most appropriate for these variables. The yield strength and wall thickness of linepipe are other examples of the type of variables that can be treated in this manner. The linepipe properties are important if using corrosion models to calculate mean time to failure. Rather than using the minimum values for each, based on the specified material and the variation allowed within the specification, typical distributions can be defined for each value. Such variables tend to be distributed normally around a mean with the specified minimum properties defining a lower bound.
Non-Linear Relationships

Many variables in corrosion rate predictions, such as the level of CO2 in the gas phase, are based on best guess or on limited well test data. No attempt is made to define the uncertainty in these data and this is a major limitation of deterministic modelling. In defining the distributions of such variables, the mean value should be based on the best guess or well test data in a similar way to the deterministic approach. However, a range of possible values should be considered. In the absence of any other information, the distribution of values is likely to be symmetrical around the mean with the greatest probability associated with values close to the mean. The Normal distribution is a familiar example of this type and should be used. It should be noted that using a symmetrical distribution, such as a Normal distibution, does not correspond to using a single value equal to the mean if the variable under consideration has a non-linear relationship with the outcome. For example, the corrosion rate prediction model used by BP states that:

36

THE USE OF PREDICTIVE MODELS DURING DESIGN

Corrosion Rate ( CO 2 partial pressure )

0.67

Therefore, the corrosion rates associated with the CO2 partial pressure values in the Normal distribution that are greater than the mean value are closer to the mean corrosion rate than those associated with the values below the mean CO2 partial pressure. In other words, defining symmetrical distributions for variables whose influence is described by a power < 1 produces a non-symmetrical distribution of outcomes (predicted corrosion rates). The mean value of this distribution will be lower than the single value calculated using the mean of the input variable. The same applies to all symmetrical distributions, including Uniform distributions. In the previous section on 'worst case design', the uncertainties regarding operating temperature and pressure were discussed. In both cases, Uniform Distributions were used to define the range of possible values. In corrosion rate modelling, both these inputs have non-linear relationships with the outcome (predicted corrosion rate). The effect of pressure is moderated by a fugacity coefficient related to the non-ideality of CO2. Therefore, considering a range of pressures distributed symmetrically around a mean value will tend to reduce the predicted corrosion rate. The effect of temperature on predicted corrosion rates is strongly non-linear. At higher temperatures, the role of protective corrosion products or scales can be important. There is a great deal of uncertainty in the effects of these scales but the bounds of the expected values can be defined using existing models. One approach would be to use a log normal distribution, defined as follows: 1. The de Waard & Milliams unscaled rate (upper bound), 2. The de Waard & Milliams fully scaled rate (lower bound), 3. A modal value equivalent to the standard BP approach that uses the scaling temperature to calculate the corrosion rate for all temperatures above this. Again, the outcome of considering a range of temperatures symmetrically distributed around a mean will tend to be a lower corrosion rate estimation than found by calculating a single value at the mean temperature.
Summary of Inputs to a Monte Carlo Analysis

Each input into a corrosion rate prediction should be considered and a range of possible outcomes defined. By consideration of the way in which the value may vary in practice, a distribution function can also be defined. This may have to be done subjectively but the following basic rules offer some guidance. In the following examples, distributions are shown that have been used in the Crystal Ball software.

37

THE USE OF PREDICTIVE MODELS DURING DESIGN

1. Where variations would be due to nature, such as the difference in CO2 levels around the field, a Normal Distribution should be used with a mean equivalent to the best guess. Figure 7 shows an example of a Normal Distribution describing the expected variation in CO2 levels, centred around a mean of 5%.

Figure 7: An Example of a Normal Distribution for the concentration of CO 2 in a gas. The Mean Value is 5 mole% with a range of 3 to 7 mole%.

38

THE USE OF PREDICTIVE MODELS DURING DESIGN

2. Where an input may vary over a wide range but would be expected to be skewed around the 'best guess' or predicted value, a Log Normal Distribution should be used. The effects of high temperature scaling would be an example of this type of distribution, or the pit depth at which inhibitors fail to control corrosion. Figure 8 shows the Log Normal Distribution used to describe the critical pit depth with a modal value of 8 mm and a range of 5 to 12mm.
Figure 8: An Example of a Log Normal Distribution describing the critical pit depth.

39

THE USE OF PREDICTIVE MODELS DURING DESIGN

3. Where a value may occur equally often within the defined range e.g flowline operating pressure, a Uniform Distribution should be used, i.e. all values are equally likely to occur. Figure 9 shows how a range of flowline operating pressures can be described. In this case the range of 1,000 to 1,200 psi has been used.
Figure 9: An Example of a Uniform Distribution Describing the Flowline Operating Pressure

Table 9 summarises the assumptions used in a recent probabilistic study into mean time to failure, based on CO2 corrosion risks. As the study looked at failure mechanisms as well as corrosion rates, some of the factors apply to the linepipe steel while others apply to the CO2 prediction model. The 'Standard Value' corresponds to the value that would be used in a deterministic study. The Table does not attempt to fully define the distributions in a statistical sense but more information is available from the authors if required.

40

THE USE OF PREDICTIVE MODELS DURING DESIGN

Table 9: Summary of Variables Modelled, the Values that would be Assigned Using a Standard Approach, and the Range of Values Used in the Example Study

Component in study

Variable

'Standard Value'

Range Used

Distribution

Linepipe Linepipe Linepipe Fluids

Wall thickness Yield Stress Flow Stress CO2 Content

Mean = 0.75" SD = 0.01 SMYS Mean = 70 ksi e.g. 65 ksi SD = 2.5 ksi ---1.15 x Yield Stress 5 mole% Mean = 5% SD = 0.72 85 - 110 oC 1,000 - 1,200 psi Cormed * 0.25 units Unscaled to fully scaled 65 - 95% 5 - 12 mm 0 - 90%

e.g. 0.75"

Normal Normal Normal Normal

Fluids Fluids

Temperature Pressure

110oC 1,200 psi

Uniform Uniform Normal Log Normal

Corrosion Water pH Cormed * model prediction Corrosion Corrosion rate >Rate at scaling model scaling T oC temperature Inhibitor efficiency Inhibitor efficiency Inhibitor efficiency Inhibitor availability Critical pit depth Inhib. effic. > critical pit depth 90% 8 mm 0%

Log Normal Log Normal Uniform

Note * Cormed is a software programme which can predict in-situ pH values of oilfield brines. Figure 10 shows the output from a Monte Carlo simulation, using 20,000 iterations to determine the distribution in outcomes (predicted corrosion rate) due to the variation in inputs detailed above. The most likely corrosion rate is circa 1 mm/yr. While there is a possibility that higher or lower rates occur, the probability of such rates decreases the further they are from the most likely outcome.

41

THE USE OF PREDICTIVE MODELS DURING DESIGN

Figure 10: Typical Outcome of the BP Corrosion Rate Model Run Using a Probabilistic Approach

Forecast: Predicted Corrosion Rates 20,000 Trials


.113

Frequency Chart

313 Outliers
2260

.085

.057

.028

565

.000 0.00 1.13 2.25 mm/yr 3.38 4.50

Effect of Corrosion Inhibitors This section represents a significant shift from recommendations and therefore is covered in some detail.
Applicability of the Guidelines

previous

BP

The guidelines on the reliance to be placed on corrosion inhibitors presented here have been based on experience gained with continuous injection systems. The success of batch treatments with corrosion inhibitor is less well documented and generally this approach to corrosion control is less reliable. These guidelines should therefore not be used when designing systems that will be protected with batch treatments - this effectively rules out their use for the majority of downhole applications. Instead, it is recommended that relevant operational experience with batch treatments is sought before designing on the basis of batch inhibition. The authors will be able to assist in sourcing relevant operational experience. Previous BP guidelines have dealt with the affect of corrosion inhibitors on CO2 corrosion by assigning a corrosion inhibitor efficiency. This described the extent to which an inhibitor reduced the predicted rate and a figure of 85% was originally used, later raised to 90%. This was despite laboratory observations that showed inhibitors could reduce corrosion rates by 95% or more. However, it was accepted that in the field, inhibitor is not delivered at the recommended dose rate for 100% of the time and therefore a degree of conservatism is necessary when estimating the benefits of inhibitors.

Inhibited Corrosion Rates

42

THE USE OF PREDICTIVE MODELS DURING DESIGN

One major limitation with inhibitor efficiencies is that it allows no consideration of the effects due to increased dose rates or the development of better chemicals. It is well known that increasing the dose rates of corrosion inhibitors up to a certain level reduces the corrosion rate. Figure 11 shows the relationship between dose rate of inhibitor and corrosion rate on corrosion coupons at Prudhoe Bay. Clearly, the inhibitor efficiency is not a constant value and increasing the inhibitor concentration (or changing the chemical for a more efficient one) enables lower corrosion rates to be achieved.
Figure 11: The Improvement in Performance of a Corrosion Inhibitor with Increasing Concentration
100

1/corrosion rate (years per mm)

10

1 40 50 60 70 80 90 100 110 120 130 140

Corrosion Inhibitor Concentration - ppm

A second major limitation with using a single value for corrosion inhibitor efficiencies is that they are unlikely to be constant across the whole range of field conditions. CO2 corrosion models can handle input values across a wide range and moderation factors have been developed over the years to reduce the conservatism due to the extrapolation of the data set used to develop the model. However, no such moderation factors have been developed for corrosion inhibitor efficiencies and by applying a blanket efficiency, it is assumed they are constant across the range of applications. BP is fortunate in having one of the more corrosive fields in Prudhoe Bay. This field also lends itself to effective corrosion monitoring due to the use of above-ground flowlines and there is a great deal of data on inhibited corrosion rates. There is a good relationship between observed corrosion rate and inhibitor concentration, as shown in Figure 12. In this Figure, the effect of the increased dose rate of chemical between January 1994 and September 1996 can be seen in the increased efficiency of the chemical, based on the predicted corrosion rates using BPs CO2 corrosion rate prediction model.

43

THE USE OF PREDICTIVE MODELS DURING DESIGN

Average Corrosion Inhibitor Concentration - ppm

99.40% 100

99.20% 80 99.00% 60 98.80%

40 98.60%

20

98.40%

0 Jan-94 May-94 Sep-94 Jan-95 May-95 Sep-95 Jan-96 May-96 Sep-96

98.20%

Date

In Figure 12 all efficiency values lie within the range 98.6% and 99.7%, apparently extremely good performance but in January 1994 only 40% of the flowlines at PBU had acceptable rates of corrosion, defined as corrosion rates under 2 mpy (0.05 mm/yr.) based on corrosion probes - see Figure 13. The improvement in performance from January 1994 onwards correlates with the increase in average dose rates shown in Figure 12.

44

'Average' Corrosion Inhibitor Efficiency

Figure 12: The Relationship Between Corrosion Inhibitor Dose Rate and Observed Efficiency at Prudhoe Bay

140

99.80%

Corrosion inhibitor concentration 120 99.60% Corrosion inhibitor efficiency, defined using BP's model

THE USE OF PREDICTIVE MODELS DURING DESIGN

Figure 13: Historical Record of Corrosion Rates in PBU Flowlines Showing Improving Performance Since January 1994

Percentage of Production Lines with Corrosion Under Control


100% 80% 60% 40% 20% 0% -20% -40% -60% -80% -100% Jan-90 Jan-91 Jan-92 Jan-93 Jan-94 Jan-95 Jan-96 Note Covers 3 phase production lines >6" in diameter with WLCs including LDFs, LP, HP and GHX.

2 < CR < 5 1 < CR < 2 < 2 mpy by Qtr

CR >5 mpy CR < 1 mpy

Prudhoe Bay was constructed before the development of the earlier BP guidelines on CO2 corrosion, but if their flowlines were to be constructed today using the same materials and corrosion allowances, it would infer a corrosion inhibitor efficiency of approximately 98%. As PBU have now demonstrated that corrosion control of their system is possible it is clear that inhibitors can be effective under highly corrosive conditions. This in turn indicates that either:

r Higher inhibitor efficiencies can be assumed in more aggressive conditions, or r Corrosion inhibitor efficiencies are not the correct way to describe the role of inhibitors in corrosive service.
The former premise does not lend itself to design as it would require a sliding scale of inhibitor efficiencies and the field data is not available to allow this to be produced. The latter is the belief of several oil companies who do not use inhibitor efficiencies, preferring to use a design corrosion rate for inhibited systems in the range 0.1 to 0.3 mm/year. For mildly corrosive conditions (~1.0mm/year) the use of an efficiency of 90% generally works well. However, for highly corrosive conditions (~10mm/year) it would result in a conservative estimate of the inhibited corrosion rate. This adds weight to the argument that the role of corrosion inhibitors can not be described by efficiencies.

45

THE USE OF PREDICTIVE MODELS DURING DESIGN

BPs data indicate that inhibited corrosion rates of 0.1 mm/year are possible under optimum conditions of high inhibitor dose rates and optimised chemicals. This is confirmed with inspection data from PBU where flowlines which have been effectively inhibited have pipewall corrosion rates of less than 0.1 mm/yr.
Applications Where Inhibitors Are Less Than Fully Effective

In general, inhibitors require free and regular access to the steel surface to be effective. Anything that interferes with this will reduce their effectiveness to low or negligible levels. Examples of low or stagnant flow situations are vessels, instrument and drain piping and tanks. Historically, inhibitors have not been assumed to work well in these environments and other corrosion control measures are used, such as coatings and/or cathodic protection. Inhibitors also perform poorly in low velocity pipework and pipelines, particularly if the fluids contain solids such as wax, scale or sand. Under such circumstances, deposits inevitably form at the 6 oclock position, preventing transportation of the inhibitor to the metal surface. Flow velocities below approximately 1.0 m/s should be avoided if inhibitors are to provide satisfactory protection and this will be critical in lines containing solids. The figure of 1.0 m/s is a rule-of-thumb which has been used in the industry for many years. However, it is now possible to calculate the velocity more accurately, using an approach developed by the 'Corrosion in Multiphase Systems Centre' at Ohio University [18]. The work agrees with the rule of thumb for most black oil systems but allows more accurate quantification if the minimum velocity is restrictive.

Operating Costs Associated With Corrosion Inhibition

The costs associated with corrosion inhibition are driven by the volume of chemical used per annum and the chemical cost. There may be some incidental costs associated with the provision and maintenance of injection equipment but increasingly this is being handled by the chemical suppliers and is therefore covered by the chemical cost. In general, inhibitors are most attractive when protecting long lengths of pipeline while they are rarely cost effective when protecting short runs of process piping. The dose rates required are dependent on factors such as liquid throughput, CO2 partial pressure, pH and flow regime. Dose rates are not dependent on the length of pipeline or pipework being treated and

46

THE USE OF PREDICTIVE MODELS DURING DESIGN

therefore the same operating cost is incurred in protecting 10 metres of pipework as is required to protect 20 km of flowline. Corrosion resistant materials are likely to offer lower life cycle costs for pipework while carbon steel plus inhibition tends to be the cheapest method of constructing and operating flowlines [19].
Table 10: Dose Rates of Corrosion Inhibitors into Several North Sea Export Pipelines, Based on Total Fluid Volumes

Field Beatrice Brae Bruce * Forties Pipeline * Magnus Miller * Nelson Enterprise * Scott Amerada Hess * AVERAGE

Dose Rate (ppm) 40 10 46 26 20 35 17 9 25

Note * - These fields deploy concentrated corrosion inhibitors to improve logistics offshore. The quoted dose rates correspond to the standard product, manufactured by the same supplier. At Prudhoe Bay the field-wide average corrosion inhibitor injection rate is 110 ppm, with maximum rates of 250 ppm in certain flowlines, based on water production (typical water cuts are 50%). These rates reflect the rapid corrosion experienced in some PBU flowlines in recent years. The determination of dosage rates in gas systems is not as straightforward as for liquid filled lines. The three methods which are commonly used to do this are: 1. Based on Gas Flow. This is the most commonly used method and a common rule of thumb is to apply 1 pint of inhibitor to every 1 million standard cubic feet of gas (1 pint/MMscf). Actual values are found to vary enormously in the range of 2 and 0.05 pints/MMscf of gas. 2. Based on the Water Content in the Pipe Line. This is the method favoured by corrosion engineers as it usually indicates a very low requirement for inhibitor. It is common to assume a dosage of 200 ppm

47

THE USE OF PREDICTIVE MODELS DURING DESIGN

of chemical in the water. This method will often give erroneously low values, especially when the water content is very low and/or the pipeline is very long. This is because the volume predicted will be too low to allow a film to be build up over the entire surface of the pipe. 3. Based on the Formation of a Protective Film. This is probably the least used method but one whch provides a good check on the values obtained from the first two methods. Typically it is the volume required to form a 0.05mil (1 micron) film over the entire internal surface of the pipe. This volume is then applied continuously on a daily basis. If the product is to be applied as a batch treatment the volume is increased by a factor of ten (x10). In practice it is sensible to do all three calculations and to use the greatest volume as the starting point. This should hopefully be the most conservative volume required. Again, highly corrosive duties associated with high temperatures or CO2 partial pressures will tend to require dose rates towards the upper end of this scale. Chemical costs vary from supplier to supplier and may be tied in with the provision of other services such as corrosion monitoring. However, for the purposes of life cycle costing a chemical cost of US$8 per US gallon is reasonable. On this basis, corrosion inhibitor costs 0.84 cents to 8.4 cents per barrel at inhibitor dose rates of 25 to 250 ppm. There will also be costs associated with monitoring and inspection. These aspects are beyond the scope of this document but are covered in detail in SELECTING MATERIALS FOR WEALTH CREATION: A Material Selection Philosophy Based On Life Cycle Costs [17]. Predicting the Effectiveness of Corrosion Inhibitors - The Inhibitor Availability Model Due to the limitations of corrosion inhibitor efficiencies as a design tool, the inhibitor availability model has been adopted. This approach can be used to define a corrosion allowance as follows: CAtotal = CAinhibited (x years @ 0.1 mm/yr.) + CAuninhibited (y years @ uninhibited rate) This approach assumes that the inhibited corrosion rate is unrelated to the uninhibited corrosivity of the system and all systems can be inhibited to 0.1 mm/year. The approach also acknowledges that corrosion inhibitor is not available 100% of the time and therefore corrosion will proceed at the uninhibited rate for some periods.

48

THE USE OF PREDICTIVE MODELS DURING DESIGN

In the context of this model, corrosion inhibitor availability infers the presence of a suitable corrosion inhibitor at sufficient concentration to reduce the corrosion rate to 0.1 mm/yr. The factors that lead to inhibitor availability below 100% are: Inhibitor injection equipment is not available on Day 1 of operations. Injection equipment requires maintenance and repairs. Operators set the dose rate incorrectly. Chemical is not available when required. Chemical dose rate is less than optimum. This can be due to a variety of reasons including lack of response to increases in throughput, or water cut or sand rate. r Well stimulation fluids such as hydrochloric acid are produced along with the crude oil and reduce corrosion inhibitor effectiveness. r The corrosion inhibitor injection facilities are used for delivery of other oilfield chemicals such as demulsifiers or combined products such as scale and corrosion inhibitors. r Inhibitors are deployed via large bore pipework (instead of via injection quills) and are not dispersed in the flow stream for some distance, providing poor protection. All of these factors and others not listed have lead to less than optimal delivery of corrosion inhibitor into production equipment in BPX. No asset is immune to such problems and therefore the maximum inhibitor availability that should be assumed is 95%. In many instances, a lower availability should be assumed; see, 'Recommended Values For Use in the Inhibitor Availability Model, pp 51.' Words of Caution Production data from Cusiana shows that their 12 inhibitor injection skids averaged 99.2 % availability over the second half of 1996, an identical figure to that generated at a new gas treatment plant in the Middle East. This is probably close to the maximum that inhibitor injection units can be available, bearing in mind the requirements for chemical feedstock, power and the reliability of the pumps. However, this should not be used as a basis for assuming an inhibitor availability of greater than 95%. Figure 14 shows the delivery of corrosion inhibitor against the target rate for a North Sea platform. There was only one instance when the inhibitor injection system was not delivering chemical - during March 1993 - but there were also only 3 short periods where the chemical was fully available with respect to the target dose rate.

r r r r r

49

THE USE OF PREDICTIVE MODELS DURING DESIGN

Figure 14: The Availability of Corrosion Inhibitor into a Main-Oil-Line over an 18 Month Period

100

80

60

Target = 50ppm

40

20

0 January 1993

March 1993

May 1993

July 1993

September November 1993 1993

January 1994

March 1994

May 1994

At the project stage, it is difficult to determine the availability of inhibitor in future years but relatively easy to ensure inhibitor is available on day one. The provision of chemical injection equipment is often outside the scope of EPIC contracts and therefore assets are brought on-stream without the necessary facilities to inhibit valuable equipment. In previous projects, this has taken up to 2 years to correct and therefore the best inhibitor availability that can be achieved will be 90%, assuming a 20 year design life. If the provision of chemical injection equipment is brought inside the scope of the EPIC contract, measures can be taken to ensure inhibitor is available on day 1 of operations. Achieving good inhibitor availability during operations is partly down to system design and partly due to management of the changing corrosion risk. Inhibitor injection systems are simple systems and lend themselves to high levels of mechanical availability. This can be improved further through the use of low level warning devices on the storage tanks and dose rate gauges such as the sight glass or more complicated dose rate monitoring systems. Together, these two simple measures will help to ensure that the target dose rate is achieved for a high proportion of the time. Ensuring the target dose rate is correct is more difficult and requires that constant changes to the target are made to reflect changes in production rate, water cut etc In extreme cases, this may require weekly tailoring of the target dose rate. This is where corrosion control programmes can fail and therefore it is important that the materials or corrosion engineer concentrates on this aspect.

50

THE USE OF PREDICTIVE MODELS DURING DESIGN

Figure 15 shows the feedback loop that is required for effective management of corrosion using chemicals. As chemical inhibition is the only viable method for controlling internal corrosion, it is important that the deployment of chemical receives attention.
Figure 15: The Feedback Loop that Must be in Place for Corrosion Control to Work Effectively

Corrosion Models Field experience

Experience from other assets

Quantify Risk
CorrOcean FSM UT mats

Apply Controls

Monitor Effectiveness
Corrosion probes

Chemical inhibition

Intelligent pig inspections

Recommended Values for Use in the Inhibitor Availability Model The degree to which a project or asset can rely on corrosion inhibition will depend heavily on the investment made to ensure satisfactory operation of the feedback loop in Figure 15. The different approaches to managing this feedback loop enable five categories to be defined which in turn allow recommendations to be made on the values used for inhibitor availability. In all cases, it is recommended that the inhibited corrosion rate is assumed to be 0.1 mm/yr. The inhibitor availability value will reflect the approach of an asset to corrosion inhibition. The following categories have been defined to cover the entire range, based on predicted corrosion rates. Each asset or project may have equipment corresponding to two or more categories, as the modelled corrosion rate will vary throughout the facilities. The categories are summarised below and discussed in detail in the following sections, starting with the lowest corrosion risk.

51

THE USE OF PREDICTIVE MODELS DURING DESIGN

r Category 1 - Benign fluids where corrosion inhibitor usage is not anticipated. Predicted metal losses should be accommodated by corrosion allowance alone. r Category 2 - Corrosion inhibitor will probably be required but at the predicted corrosion rates there will be sufficient time to review the need for inhibition based on inspection data. r Category 3 - Corrosion inhibition will be required for the majority of field life but the facilities will not be available from Day 1, limiting the maximum effectiveness of a corrosion control programme. r Category 4 - Corrosion inhibition is relied on heavily and will be required for the entire period of operation. Inhibitor must be available on Day 1 to ensure maximum probability of success for the corrosion control programme. r Category 5 - Carbon steel and corrosion allowance with corrosion inhibition is unlikely to provide integrity for the full field life, thereby requiring repairs or replacements. Should only be considered once environmental and economic analyses have shown this to be more cost effective than using corrosion resistant materials - an option of last resort.
Categories 2 and 4 are illustrated schematically in Figure 16. Categories 1, 3 and 5 can be considered in a similar manner.

52

THE USE OF PREDICTIVE MODELS DURING DESIGN

Figure 16: The Concept of Inhibitor Availability in Relation to Consumption of Corrosion Allowances

Spare CA

SAFE UNSAFE

Derating, repair or replacement required in Year 10

Cross section of pipe or vessel on Day 1

Field Life (years)

20

Category 4 - red example: Uninhibited corrosion continues at high rate for 2 years, when inhibition is started. However, the inhibitor is incapable reducing the corrosion to a sufficient degree and de-rating or replacement will be required at Year 10. In this case, 18 years of inhibition (equivalent to 90% availability) is not sufficient due to the high rates of uninhibited corrosion in Years 1 and 2. The availability of inhibitor must be improved to 95% if carbon steel and corrosion inhibition is to work satisfactorily and therefore the system should be designated as a Category 4 and designed and operated accordingly. Category 2 - blue example: Uninhibited corrosion proceeds at a moderate rate for 10 years, when inhibition is started. The inhibited rate is low enough to enable full field life to be reached with corrosion allowance to spare. In this case 10 years of inhibition, equivalent to 50% availability is satisfactory. This would place this example in Category 2 as there is ample time to detect corrosion prior to the implementation of a corrosion control programme.

53

THE USE OF PREDICTIVE MODELS DURING DESIGN

In these examples, once inhibition is initiated in Year 2 or 10, it is shown as being effective at controlling corrosion at 0.1 mm/yr. for the remaining period i.e. 100% availability for the remaining period. In practice, this will not be the case and inhibitor availability will be less than 100% due to the reasons described pp 49. This would see the lines representing the loss of corrosion allowance becoming step shaped, corresponding to the periods of inhibitor availability and non-availability. Figure 17 provides a pictorial representation of these relationships.
Figure 17: A Pictorial Representation of the Relationship between Corrosion Rates, Design Life, Inhibitor Availability and Corrosion Allowance.
Knowns Uninhibited corrosion rate - from model Inhibited corrosion rate = 0.1 mm/yr. Design life e.g. 20 years Variables Inhibitor Availability Corrosion Allowance

In general, decreasing CA: Reduces CAPEX Increases OPEX Increases monitoring

Options: Increase CA, decrease availability Decrease CA, increase availability

Risk category determines requirements for: Corrosion control Monitoring Inspection

Outcome Corrosion Risk Category 1 to 5

Velocity limitations relate to inhibited fluids

Fluid Velocity C-factor < 100, no change C-factor 100-135, + 1 category

Table 11 shows some examples of how the corrosion risk category is determined.

54

THE USE OF PREDICTIVE MODELS DURING DESIGN

Table 11: Some Examples of how the Corrosion Risk Category is Determined. Knowns:

Worked example for determining optimum corrosion risk category


Variables: Inhibitor availability = 0 to 95% Corrosion allowance = 0 to 8.0 mm

Uninhibited corrosion rate = 2.0 mm/yr. Inhibited corrosion rate = 0.1 mm/yr Design life = 20 years

Design as Category 1 System Inhibitor availability = zero Corrosion allowance required: (20 x 2.0) + (0 x 0.1) = 40 mm Not a practical option: corrosion allowance > 8.0 mm Design as Category 2 System Inhibitor availability = 49% Corrosion allowance required: (10 x 2.0) + (10 x 0.1) = 21 mm Not a practical option: corrosion allowance > 8.0 mm Design as Category 3 System Inhibitor availability = 90% Corrosion allowance required: (2 x 2.0) + (18 x 0.1) = 5.8 mm Practical option: moderate corrosion allowance and corrosion control, monitoring and inspection requirements Design as Category 4 System Inhibitor availability = 95% Corrosion allowance required: (1 x 2.0) + (19 x 0.1) = 3.9 mm Practical option: minimal corrosion allowance with requirements for elaborate corrosion control, monitoring and inspection requirements In this example, the choice is between designing as a Category 3 or 4 system. Both are practical solutions and the optimum balance for a project will be determined by the relative cost of the extra 1.9 mm corrosion allowance required for a Category 3 system compared with the additional costs of the control, monitoring and inspection incurred with a Category 4 system. In general, long pipelines will be more cost effective when designed to a higher category while shorter pipelines or process piping will be more cost effective as a lower category system.

The workbook provided on the disc with these guidelines contains a spreadsheet for determining the corrosion risk category of a given system.

55

THE USE OF PREDICTIVE MODELS DURING DESIGN

Designing and Operating a Category 1 Corrosion Control System

Category 1 - Basis of Design Assumed inhibitor availability = 0% Maximum tolerable uninhibited corrosion rate = 0.4 mm/yr. This approach will be valid for applications where the predicted cumulative corrosion rate over field life can be accommodated by a corrosion allowance. In practice, this means a maximum predicted corrosion rate of 0.4 mm/yr., assuming a design life of 20 years and a maximum corrosion allowance of 8 mm. Longer or shorter design lives will change this rate accordingly. Corrosion inhibition provides a fallback measure in case the actual corrosion rate are higher than predicted due to changes in field conditions or unforeseen circumstances. Category 1 - Corrosion Monitoring and Inspection Requirements The fluids must by definition be benign and corrosion rates low. Corrosion monitoring equipment such as corrosion probes and coupons will respond slowly to changes in corrosion rates and will be of little practical benefit. Detection of unexpectedly high corrosion rates remains important as the insitu corrosion rates may be higher than predicted. However, rates are unlikely to exceed the predicted rate by more than a factor of 2 (i.e. 0.8 mm/yr. maximum) and therefore the inspection programme will be capable of detecting such attack. This can provide an early warning system, allowing time for implementation of a corrosion control programme if required. The usual requirements of an inspection programme apply. In particular, it should anticipate localised corrosion at areas such as welds and the 6 oclock position of low flow rate lines. Category 1 - Corrosion Control System Requirements As the design of the facilities does not rely on the use of corrosion inhibition, there is no requirement to incorporate corrosion injection facilities into the design.

Designing and Operating a Category 2 Corrosion Control System

Category 2 - Basis of Design Assumed inhibitor availability = 50% Maximum tolerable uninhibited corrosion rate = 0.7 mm/yr.

56

THE USE OF PREDICTIVE MODELS DURING DESIGN

Category 2 equates to mildly corrosive fluids where the predicted corrosion rate is too high to be accommodated by corrosion allowance alone but where corrosion inhibition should not be required for the full field life. In practice, this approach is only valid for predicted corrosion rates of up to 0.7 mm/yr., again assuming an 8 mm corrosion allowance and 20 design life. Using a corrosion inhibitor efficiency of 50% infers that approximately 9 years of uninhibited corrosion can be accommodated before 95% reliance on inhibition must be assumed for the remaining 11 years of a 20 year field life. This provides time for corrosion to be detected via inspection programmes. Category 2 - Corrosion Monitoring and Inspection Requirements A design of this type relies heavily on monitoring systems to detect the onset of corrosion at a rate requiring inhibition. This will require monitoring of process changes such as temperature, flow velocity and water cut. Direct corrosion rate monitoring will also be required. However, due to the relatively low corrosivity of fluids, response from corrosion probes and coupons may be poor. Due to the relatively low corrosivities of the fluids, inspection programmes will also play a vital role in detecting the onset of corrosion. Uninhibited corrosion losses of half the corrosion allowance over a 3 to 5 year period will be detectable by inspection techniques and will still enable corrosion inhibition to reduce rates to acceptable levels over the remaining field life. Selecting corrosion allowances using the BP model will ensure several years of corrosion can be accommodated prior to inhibition being required. Category 2 - Corrosion Control System Requirements The corrosion control system must be capable of being commissioned and to begin injection as soon as changes in the corrosion rate are detected. This means that the plant should be designed for inhibitor injection without recourse to a shutdown. In practice, this will mean that access fittings should be installed to allow fitment of corrosion inhibitor injection quills at system pressure. Provision of equipment upstream of the quill such as the pipework, dosing pumps and storage tanks can be delayed until monitoring or inspection data show that corrosion inhibition is required.

57

THE USE OF PREDICTIVE MODELS DURING DESIGN

Designing and Operating a Category 3 Corrosion Control System

Category 3 - Basis of Design Assumed inhibitor availability = 90% Maximum tolerable uninhibited corrosion rate = 3 mm/yr. This category equates to projects or assets that require corrosion inhibition for almost the full life of the field but do not include the specification and provision of corrosion control and monitoring facilities into the overall project scope for use on Day 1 of operations. In practice, this may mean that corrosion control equipment is not on site and commissioned for 12 months or more and therefore the reliance that can be placed on inhibition is less than 95%. A delay of 12 months means that corrosion inhibitor availability must average 95% over the remaining 19 years to achieve an overall availability of 90%. This limits the maximum predicted corrosion rate that can be successfully accommodated to 3.1 mm/yr., assuming a corrosion allowance of 8 mm and a 20 year design life. Category 3 - Corrosion Monitoring and Inspection Requirements A facility in this category will have a predicted corrosion rate of 0.7 to 3.1 mm/yr. Failure of the corrosion control programme can lead to failure in under 3 years if the corrosion allowance is selected in accordance with the guidelines. Reliance on the corrosion control programme is therefore high, particularly as it will not be present on Day 1 of operations. The corrosion monitoring system must be capable of detecting changes in corrosion rates within weeks if the target rate of inhibitor injection is to be constantly revised to ensure the overall availability of 90% is achieved. The recommended techniques that are capable of providing such resolution are ultrasonic mats and the CorrOcean FSM. Category 3 - Corrosion Control System Requirements It is recognised that the corrosion control system will not be available on Day 1 of operations. However, it must be capable of being commissioned without recourse to a shutdown. In practice, this will mean that access fittings should be installed to allow fitment of corrosion inhibitor injection quills at system pressure. The corrosion inhibitor should have been preselected and the initial dose rate should be based on either laboratory trials or similar operating experience elsewhere. Provision of equipment upstream of the quill such as the pipework, dosing pumps and storage tanks should also be planned during the design phase to unsure there is adequate deck space and power supplies to enable the

58

THE USE OF PREDICTIVE MODELS DURING DESIGN

system to be commissioned quickly once it arrives. The system should incorporate dose rate meters and low level warning devices on the storage tank.
Designing and Operating a Category 4 Corrosion Control System

Category 4 - Basis of Design Assumed inhibitor availability = 95% Maximum tolerable uninhibited corrosion rate = 6 mm/yr. This category applies to equipment that require corrosion inhibition to be present for the full design life of the field to ensure satisfactory integrity from carbon steel equipment. The reliance on corrosion inhibition is high and a failure could occur in a little over 1 year if the corrosion control programme fails. To achieve inhibitor availability of 95%, the corrosion control system must be operational on Day 1. To ensure this happens, it is recommended that the provision of the control system is brought within the scope of the overall project. Category 4 - Corrosion Monitoring and Inspection Requirements A facility in this category will be handling highly corrosive fluids and the corrosion control programme will require constant optimisation to ensure the corrosion allowance is not consumed prematurely. This may require dose rates of chemicals to be checked on a weekly basis and the sensitivity of corrosion monitoring devices must reflect this. The recommended techniques that are capable of providing such resolution are ultrasonic mats and the CorrOcean FSM. Category 4 - Corrosion Control System Requirements The corrosion control system must be commissioned and working on Day 1 of production. The corrosion inhibitor should have been pre-selected and the initial dose rate should be based on either laboratory trials or similar operating experience elsewhere. The system should incorporate dose rate meters and low level warning devices on the storage tank.

Designing and Operating a Category 5 Corrosion Control System

Category 5 - Basis of Design Assumed inhibitor availability > 95% Uninhibited corrosion rate > 6.0 mm/yr. This category of corrosion risk is beyond BPs recommended practice. Predicted corrosion rates beyond 6 mm/yr should not generally be handled

59

THE USE OF PREDICTIVE MODELS DURING DESIGN

through a combination of carbon steel with corrosion allowance and corrosion inhibition. Instead, corrosion resistant materials should be considered. There will always be specific cases where corrosion resistant materials are not feasible or where previous operating experience indicates that carbon steel will corrode at a lower rate than indicated by the model. However, the risks involved in operating such a system are high and repairs or replacement of equipment should be expected during the field life. This is unlikely to be cost effective when lost production costs and potential environmental damage are considered and these areas must be addressed if such highly corrosive fluids are to be handled or transported using carbon steel. Category 5 - Corrosion Monitoring and Inspection Requirements Assuming the technical, environmental and financial factors of operating a carbon steel facility of this type have been considered and answered satisfactorily, the monitoring requirements will be similar to those for a Category 4 system. Category 5 - Corrosion Control System Requirements Assuming the technical, environmental and financial factors of operating a carbon steel facility of this type have been considered and answered satisfactorily, the control system requirements will be similar to those for a Category 4 system. Table 12 summarises the recommendations made in respect of each category. Table 12 also classifies when an intelligent pig inspection should be carried out for the various corrosion risk categories. These classifications are described on page 52. These can be scheduled by a variety of means, depending on the amount of information available for the system. If there is extensive process and corrosion monitoring data together with extensive operational experience of the system, it may be possible to schedule inspections based on gathered data i.e. using ER probe data as a trigger. However, until experience and confidence are gathered corrosion modelling offers the best method. The reliance on monitoring and inspection is greater for Categories 5, 4 and 3 than for Categories 2 and 1 and therefore inspection should occur earlier in the fields life.

60

THE USE OF PREDICTIVE MODELS DURING DESIGN

Table 12: Summary of Criteria and Requirements for Corrosion Risk Categories 0 to 4

Corrosion control system requirements


Corrosion Inhibitor Corrosion Assumed Risk availability rate (max) dose rate System availability System sophistication Scheduling 1st inspection

Monitoring requirements, based on location


On land, above Ground On land, buried Subsea

Category 1

Zero

0.4 mm/yr.

0 ppm

None required

No requirement

Routine inspection

Process monitoring Standard inspection techniques

Process monitoring Standard inspection techniques

Process monitoring Standard inspection techniques

Category 2

50%

0.7 mm/yr.

20 ppm

Should be capable of commissioning w/o plant shutShould be included in basis of design and commissioned as soon as practical

No special requirement

Routine inspection

As Category 1 plus weight loss coupon ER probes Intelligent pig run As Category 2 plus regular inspection of bends, welds etc Continual data logging for probes

As Category 1 plus weight loss coupon ER probes Intelligent pig run As Category 2 plus FSM or UT mat system Continual data logging for all monitoring devices As Category 3 plus increased inspection frequency

As Category 1 plus weight loss coupon ER probes Intelligen pig run As Category 2 plus FSM or UT mat system Continual data logging for all monitoring devices As Category 3 plus increased inspection frequency

Category 3

90%

3 mm/yr.

50 ppm

Should incorporate low level device and flow monitor in injection package

Early inspection

Category 4

95%

6 mm/yr.

100 ppm

Should be within scope of overall project and available from Day 1 Should be within scope of overall project and available from Day 1

Should include low level device and flow monitor in injection package Should include low level device and flow monitor in injection package

Early inspection

As Category 3 plus increased inspection frequency

Category 5

> 95%

>6 mm/yr.

300 ppm

Early inspection

As Category 4 plus leak detection

As Category 4 plus leak detection

As Category 4

61

THE USE OF PREDICTIVE MODELS DURING DESIGN

Comparisons of the Inhibitor Availability Model with BPs Previous Model

- The aim of the inhibitor availability model is to encompass the good track record of the inhibitor efficiency model at low to moderate corrosivities but to remove some of its conservatism in more corrosive systems. The two inputs to the model are the inhibited corrosion rate and the inhibitor availability and using different values for these can produce a whole array of outputs.
20

Recommended Corrosion Allowance for 20 year design life - mm

Figure 18: A Comparison Between the Inhibitor Efficiency and Inhibitor Availability Methods of Determining Corrosion Allowances

18 16 14 12 10 8 6 4 2 0

Inhibitor availability model based on inhibited rate of 0.1 mm/yr and availability of 95%
Corrosion allowance - efficiency method Corrosion allowance - availability method

Efficiency method based on efficiency of 90%

20.0 6.9

11.9

10.0 1.0
0.5

2.4

2.0
1

2.9

4.0
2

3.9

6.0

4.9

10

Predicted Corrosion Rate - mm/yr.

Figure 18 shows the corrosion allowance that would be recommended using the two approaches for a 20 year design life. A range of uninhibited corrosion rates are considered, from 0.5 to 10 mm/yr. which covers the range from mildly to highly corrosive fluids (less corrosive fluids would probably be handled without recourse to inhibition). In the inhibitor efficiency example, an efficiency of 90% has been assumed, in line with BPs previous practice. The inhibitor availability model uses an inhibited corrosion rate of 0.1 mm/yr. and an inhibitor availability of 95%. During the remaining 5% of the time, the uninhibited corrosion rate is used (0.5 to 10.0 mm/yr. as appropriate). Both models agree well for moderately corrosive fluids, while for mildly corrosive fluids (0.5 to 1.0 mm/yr.) the availability approach recommends a greater corrosion allowance. In practice, this may not be important as external corrosion may require a corrosion allowance of up to 2 mm and would over-ride the allowance recommended for internal corrosion.

62

THE USE OF PREDICTIVE MODELS DURING DESIGN

For highly corrosive fluids, the availability model recommends lower corrosion allowances than the efficiency model. This agrees well with the observed high efficiencies of corrosion inhibitor under highly corrosive conditions. This will increase the use of carbon steel as the standard practice is to specify carbon steel with corrosion allowances up to 8mm and to use corrosion resistant steels for more corrosive fluids. Figure 19 shows the relationship between predicted corrosion rate and the recommended corrosion allowance using the inhibitor availability method. The example shown is the same as in Figure 18 with predicted corrosion rates in the range 0.5 to 10 mm/yr. In each case, the corrosion allowance for inhibited corrosion is constant at 1.9 mm due to the assumption of an inhibited corrosion rate of 0.1 mm/yr. and the required field life of 20 years. The variation in recommended corrosion allowances is due entirely to the 5% of the time where inhibition is assumed to not occur.
12

Recommended Corrosion Allowance for 20 Year design life - mm

Figure 19: The Contribution to the Total Recommended Corrosion Allowance from the Inhibited and Uninhibited Portions of the Inhibitor Availability Model

Corrosion allowance for uninhibited corrosion Corrosion allowance for inhibited corrosion (95% availability) 10 5 0.5 1.9
0.5

10 8 6 4 2 0

1 1.9
1

2 1.9
2

3 1.9
3

1.9
5

1.9
10

Predicted Corrosion Rate - mm/yr Figure 19 helps to illustrate how important the period of uninhibited corrosion can be. In a severe case of a predicted corrosion rate of 10 mm/yr., the uninhibited period of 5% of the time accounts for 83% of the corrosion allowance. In this case, each 1% increase in the assumed availability of corrosion would reduce the total corrosion allowance by 16.6%. Table 13 gives some more details on this point.

63

THE USE OF PREDICTIVE MODELS DURING DESIGN

Table 13: The Effect of the Assumed Corrosion Inhibitor Availability on the Recommended Corrosion Allowance for a 20 year Design Life

Predicted Corrosion Rate mm/yr.. 0.5 1 2 3 5 10

CA assuming 95% inhibitor availability mm 2.4 2.9 3.9 4.9 6.9 11.9

CA assuming % reduction in 96% inhibitor corrosion allowance availability per 1% increase in mm inhibitor availability 2.3 2.7 3.5 4.3 5.9 9.9 3.3 % 6.2 % 9.7 % 11.8 % 14.2 % 16.6 %

It can be seen that highly corrosive systems must assume a high value for the inhibitor availability if carbon steel is to be used with a practical corrosion allowance. Corrosion Rates of Low Alloy Steels The corrosion rate prediction model presented here is for use with carbon steels, i.e. predominantly iron with low levels of carbon. However, some engineering materials contain a wider range of alloying elements such as chromium and nickel to improve the mechanical properties, such as strength or toughness. Such elements are commonly found in corrosion resistant materials and chromium in particular can increase the corrosion resistance of carbon steels, if present in sufficient concentration. 13% of chromium turns a carbon steel into a stainless steel, with excellent resistance to CO2 corrosion. Many claims have been made over the past 5 years of the affect of adding low levels of chromium (0.5 to 1.0%) to carbon steel. Some steel suppliers claim that 0.5%Cr can halve the CO2 corrosion rate and certainly in some tests there does appear to be a benefit. The most consistent benefit seems to be an improved resistance to mesa corrosion where large, square edged and flat bottomed pits can form. However, in other tests no benefits have been observed and it seems that the benefits may be related to microstructure rather than composition. Other researchers and oil companies have reported that inhibitors perform worse on low alloy steels than on carbon steel and therefore, in inhibited systems, there is no benefit from the addition of low levels of chromium.

64

THE USE OF PREDICTIVE MODELS DURING DESIGN

On balance, BP believe there are no proven advantages or disadvantages in terms of CO2 corrosion resistance from the presence of chromium at concentrations up to 1% in steels. It is therefore recommended that no account is taken of the presence of alloying elements at low levels and no premium should be paid for such steels. However, if the steel supplier uses low levels of chromium in the standard product, that is acceptable. Preferential Weld Corrosion Preferential weld corrosion is a problem in most systems and production systems containing CO2 are no exception. Efforts have been made to eliminate preferential weld corrosion by alloying welding consumables with various elements such as chromium, nickel and copper at low levels (circa 1%). No universal solution has been found and there are examples of either weld metal or heat affected zone (HAZ) suffering preferential attack with most welding consumables and welding procedures. The problem is not made easier by the fact that the mechanism for preferential weld corrosion is not fully understood in CO2 service. The speed of such corrosion suggests there could be a galvanic driving force. Even in benign systems where predicted rates of general corrosion are low, rates of attack at welds can be unacceptably high. This causes a problem when deciding whether a corrosion inhibitor is required for a particular application. The traditional approach has been to calculate cumulative wall losses over the life of the field using corrosion models and if the predicted wall loss is less than the available corrosion allowance, inhibitors have not been specified. However, preferential weld corrosion can proceed at rates far higher than predicted and inhibitors offer the only proven method of improving the reliability of carbon steel in such cases. There have recently been cases of preferential weld corrosion causing rapid failures in systems believed to be only mildly corrosive. Unfortunately, there can be no clear guidance for such systems but inspection programmes should recognise the risk of preferential weld attack and, if detected, corrosion inhibition should be initiated immediately.

65

THE USE OF PREDICTIVE MODELS DURING DESIGN

Effect of Pitting CO2 models are basically bare surface models with moderation factors applied to anything that affects this, such as surface scales and corrosion inhibitors. Moderation factors are used to reduce the predicted corrosion rate due to the presence of protective or semi-protective species at the surface. In other words, all such factors predict that the surface will corrode at a lower rate than would be expected if it was fully exposed to the bulk solution. Pits are one case where local corrosion rates may be higher than if the surface was exposed to the bulk solution. The environment at a corroding steel surface is different from that in the bulk due to the continual transport of reactants to the surface and products from the surface and this is reflected in the CO2 models and associated factors. These effects are generally beneficial where the corrosion process is transport controlled but can be detrimental where it is the transport of inhibitor that is limited. This can be the case in a corrosion pit where galvanic affects also play an important role. The result is that the growth rate of deep pits may accelerate. This can be seen as a loss of control by the inhibitor and may place a practical limit on the size of the corrosion allowance. For example, if an inhibitor is incapable of protecting pits deeper than 8mm, once pitting has reached this depth the corrosion rate in the pit will proceed at the uninhibited rate, i.e. 10 or 20 times faster than the bare surface rate. The increase in life due to the provision of corrosion allowance beyond 8 mm would therefore be minor. In practice, the relationship between pit depth and inhibitor efficiency is not fully understood. Field experience indicates that pits below 5 mm behave normally while pits deeper than this may corrode at a higher rate. Pitting rates up to 3 times faster than predicted have been quoted in a variety of systems. Certainly, if corrosion has reached 8 mm it is likely that the local environment within a pit will be significantly divorced from the bulk environment and hence transportation of inhibitor may be unreliable. Moreover, if corrosion has caused such metal loss, the corrosion control of the system must be poor and providing extra steel is unlikely to provide a satisfactory answer. As corrosion allowance is often consumed via pitting or localised corrosion the importance of pits should be considered when selecting the optimum corrosion allowance.

66

THE USE OF PREDICTIVE MODELS DURING DESIGN

Choosing an Optimum Corrosion Allowance The term corrosion allowance creates the impression of a uniform wastage over time leading to the gradual and controlled reduction in wall thickness. In practice, this is unlikely to be the case and the role of the corrosion allowance is to provide protection against the periods when corrosion control is poor and short term corrosion rates are high, i.e. poor inhibitor availability in the case of inhibited systems. As there is always uncertainty in the rate of corrosion (and therefore time to failure), specifying a corrosion allowance is a compromise between capital costs and reliability. Greater corrosion allowances incur greater costs but confer greater reliability. For mildly corrosive systems, low corrosion allowances of 1.5 to 3 mm are justified as they are protecting against the possibility of internal and external corrosion. In highly corrosive systems, active corrosion is almost certain to occur and therefore greater corrosion allowances should be specified to increase the mean time to failure. Some Operators specify maximum corrosion allowances and BP has tended to use the figure of 8 mm for some years. The reasons for this are: 1. Corrosion tends to be localised pitting attack and corrosion inhibitors perform poorly in deep pits. Therefore, extra corrosion allowance provides little benefit beyond approximately 8mm. 2. Carbon steel will not provide a long term solution for highly corrosive systems and if several millimetres of corrosion allowance have been lost, corrosion control of the system has not been achieved. 3. Intelligent pigs are sensitive to corrosion damage of circa 10% of wall thickness. This makes it difficult to detect the onset of corrosion in thick walled pipe which in turn means that corrosion may continue for some time before detection. It is preferable to detect corrosion early and remedy the situation and therefore thin walled pipe is preferable for detection of corrosion. 4. Welding and handling thick walled pipe is difficult and thick sections may require post weld heat treatment. Cost increases are therefore greater than the incremental increase in wall thickness. The figure of 8mm should not be seen as fixed. Each project may have different drivers in terms of the optimum balance between opex and capex costs and in certain cases, replacement of flowlines may be more

67

THE USE OF PREDICTIVE MODELS DURING DESIGN

economically attractive than high capital costs in Year 1. For one recent BPX project it was decided that localised corrosion was the main concern for the flowlines and therefore the definition of corrosion allowance should reflect this. BPs first pass defect assessment criterion for pipelines allows 20% of the pressure containing wall to be lost due to localised corrosion and the design of the corrosion allowance took this into account. This approach reduced the corrosion allowance by circa 1.5 mm and saved US$1.16 million from the cost of the flowline network. In effect, the traditional corrosion allowance was reduced from 8 mm to 6.5 mm but as the corrosion was expected to be localised, there would be 8mm of pipewall available for localised corrosion before raising any concern over integrity. In other cases, a corrosion allowance greater than 8mm may be justified but it should be recognised that the additional costs may not be reflected in the incremental increase in reliability. Use of Common Sense In specifying a corrosion allowance, the Materials Engineer should not be too pedantic. Projects often define three or more nominal corrosion allowances such as 1.5 mm, 3 mm and 8 mm. Process streams are categorised as mildly corrosive, corrosive or highly corrosive using models or experience and the appropriate corrosion allowance added to the pressure containing wall thickness defined using the appropriate code. The total required wall thickness is then reviewed against the available wall thicknesses with the next greater thickness being selected. It may be the case that the corrosion allowance just takes the total wall thickness out of one wall thickness range and into another, increasing significantly the wall thickness and the effective corrosion allowance. Example The linepipe specification API 5L lists wall thicknesses (WT) in 1.6 mm increments for 16 linepipe in the range 12.7mm to 14.3mm. If the total required WT including 6 mm corrosion allowance is 12.8mm, standard practice would be to select the 14.3 mm size. The excess 1.5 mm would add circaUS$11,500/km to the cost of the 16 flowline i.e. in excess of US$1 million for a 100km line. As the selection of the nominal corrosion

68

THE USE OF PREDICTIVE MODELS DURING DESIGN

allowance is based on imprecise models, the Materials and Pipeline Engineers should use their judgement in the selection of the final wall thickness. They may decide that a corrosion allowance of 5.9 mm is acceptable, allowing the 12.7 mm WT linepipe to be specified. Applying Models to Different Flow Regimes
Effect of Water Cut

CO2 predictive models - such as the one in this report - are based on laboratory studies, typically developed in water only systems. Various moderation factors have been applied over the years, reducing the predicted rates as experience showed them to be too conservative in their basic forms. In the approach covered here, the water cut is ignored thereby treating the pipeline or process equipment as if it was transporting 100% water. It may appear a large step to apply a model developed using laboratory data in water only systems to the field where hydrocarbons account for the majority of the throughput.
0.1 - 13 m/s 20 - 90oC 0.3 - 20 bara CO2 (0.1 m/s, 90oC, >6.5 bara CO2 excluded!!)

Figure 20: The Application of a Model Developed in Water Only Systems to Other Water-Containing Systems

Water only...

Gas / Water Oil / Water

Multiphase

However, this is not the vast over-simplification it may seem. Water wetting of the pipewall can occur at both high and low water cuts. This is despite the widely shown plot, reproduced in Figure 21 in which a relationship is proposed between water cut and corrosion rate based on water wetting. This relationship is not reliable in practise because water cuts below 1% have been known to cause rapid failures. This simply reflects the fact that

69

THE USE OF PREDICTIVE MODELS DURING DESIGN

the average corrosion rate in a system is rarely of interest: it is the maximum rate that determines time to failure. If at a water cut of 1%, 1% of the equipment is water wet 100% of the time then clearly there will be no effect of water cut on the maximum potential corrosion rate and hence time to failure.
Figure 21: An Often Presented Relationship Between Water Cut and Corrosion Rate

III
Corrosion Rate

II

I
0 0 Water Cut, % 100

Hilly terrain, changes in elevation or changes in flow direction can induce water hold-up in wells, flowlines and process equipment. Local water cuts can exceed 50% despite input water cuts of 1% or less. The water in dips may remain for weeks or months until an increase in throughput sweeps some of it out and a temporary increase in water production is seen at the outlet of the system. It is therefore unwise to rely on the formation of emulsions or similar dispersions to provide fully oil wet surfaces. It is for this reason that BP ignores the water cut in determining system corrosivity.
Effect of Flow Regime

CO2 corrosion rates are dependent on flow regime and flow velocity, hence the attempt to incorporate the effects of flow into the 1995 de Waard and Milliams model. In uninhibited corrosion, flow effects are of secondary importance, after the important controlling factors such as temperature, pressure, CO2 concentration and pH and for this reason BP have retained the earlier de Waard and Milliams model as the basis for their CO2 modelling. The 1995 model is included if the sensitivity to flow velocity changes are considered important.

70

THE USE OF PREDICTIVE MODELS DURING DESIGN

Figure 22: Different Flow Regimes Experienced at Various Combinations of Gas Flowrate and Liquid Flowrate

Liquid Flowrate

Bubble

Slug

Annular

Stratified

Stratified Wavy

Gas Flowrate

Each flow regime will cause different rates of corrosion under otherwise identical conditions and the 1995 de Waard and Milliams model offers the best method of assessing this. When considering inhibited corrosion rates under multiphase flow, the approach proposed on pp76 should be followed. In summary, velocities corresponding to C factors below 100 require no special consideration. Velocities corresponding to C factors between 100 and 135 raise the Category of the corrosion risk, e.g. from 3 to 4. Velocities corresponding to C factors greater than 135 should not be considered unless there is significant operating experience to justify this.

71

THE USE OF PREDICTIVE MODELS DURING DESIGN

Applying Models to Transportation Equipment Crude Oil Export Pipelines Crude oil transport pipelines or main oil lines (MOL) fall into two categories: 1. The fully stabilised type such as the Trans Alaskan Pipeline System and OCENSA in Colombia. 2. The partially stabilised type, such as Forties and Beatrice MOLs. The corrosivity of the fluids is different in each case and pipelines should be designed and operated accordingly.
Fully Stabilised Crude Oil Export Pipelines

In the case of fully stabilised lines, the crude oil is processed down to atmospheric pressure and may remain in tanks for some period prior to shipping. This allows water cuts to reach levels of 0.1 to 1.0%. It also allows the acid gases present in the reservoir to vent and reach very low concentrations. For example, the effective partial pressure of CO2 in an associated gas containing 2 mole% CO2 is only 0.3 psia at atmospheric pressure. The low levels of acid gases mean the potential corrosivity of the water phase will be low. Fully stabilised crude oil can therefore be considered as a non-corrosive product and typically such pipelines are constructed with minimal or zero corrosion allowance. When a corrosion allowance is specified, it is often due to concerns over external corrosion rather than internal attack. Corrosion inhibitor is not normally deployed into fully stabilised crude oil lines.

Partially Stabilised Crude Oil Export Pipelines

In the partially stabilised case, the crude oil is partially stabilised (typically offshore) and exported for final processing at a remote location (typically onshore). The crude oil in the export pipeline therefore remains corrosive as the acid gases are not vented down to negligible levels and any associated water will be corrosive. The partial pressure of gases will depend on the pressure of final processing. For example, at 7 bara the partial pressure of CO2 in an associated gas containing 2 mole% would be 0.14 bara, or 2 psia.

72

THE USE OF PREDICTIVE MODELS DURING DESIGN

Final processing pressures vary. Forties fluids are processed down to 4.5 bara while the value on Bruce is 12 bara and Brae is 16 bara. The typical range is from 1.4 bara to 20 bara and the corrosivity of the fluids will vary accordingly, along with the CO2 concentrations, temperatures etc. As the crude oil does not pass through tankage offshore, water cuts in partially stabilised lines are typically higher than in fully stabilised lines. Water cuts can reach 15% or even higher if water handling is a constraint but more typical levels are around 1%. With the removal of the majority of the CO2 and water, partially stabilised crude oil is significantly less corrosive than the non-stabilised multiphase fluids transported in flowlines, but it can not be considered as non-corrosive. The original Forties 30 and existing Beatrice export lines are adequate proof that partially stabilised crude oil is corrosive. Such pipelines should therefore be designed and operated to deal with internal corrosion. Typically a corrosion allowance of 2 to 3 mm may be specified and corrosion inhibitor should be added on a continuous basis. It is important to note that although the pressure of the oil is raised downstream of the crude oil shipping pumps, the partial pressure of CO2 does not increase. The crude oil is single phase and any remaining associated gas is in solution - see page 11. Ideally, the velocity should be maintained above 1 m/s - see page 46. Natural Gas Pipelines
Dry Gas Pipelines

To minimise or eliminate the risk of corrosion in gas pipelines it has been (and still is) common practice to dry it prior to transportation. The two most common methods involve either contacting the wet gas with dry glycol or passing it through molecular sieves. The target water content of the 'dried' gas is usually 2lbs of water for every million standard cubic feed of gas (2lbs/MMscf). However, both methods have their problems as shown by Figure 23 which shows the water content of a dry gas downstream of the glycol contactors.

73

THE USE OF PREDICTIVE MODELS DURING DESIGN

Probability

Figure 23: The Actual Water Content of a 'Dry Gas' Downstream of the Glycol Contactors Over a 2 Year Period. The Design Specification was 2lbs/MMscf.

0.25

Specification 2 lbs/mmscf
0.20

0.15

0.10

0.05

0.00 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

Water Content - lbs/mmscf

The data was gathered from a BP Asset over a two year period and it is clear that the target value of 2lbs/MMscf was rarely achieved. Thus some care should be taken when relying on the drying of gas for corrosion control and each system should be considered on a case by case basis.
Wet Gas Pipelines

As part of the drive to minimum offshore processing, gas transportation lines are increasingly being designed to operate wet i.e. the gas either enters the pipeline below its water dew-point or will drop below this temperature at some location along the pipeline. Once free water is present, corrosion becomes a concern and this must be taken into account during the design and operational phases of the pipelines life. The severity of corrosion and the potential means for controlling it depend on the operating scenario and flow regimes. If a wet gas pipeline is not going to be treated with a recycled hydrate inhibitor, corrosion inhibition is the only practical corrosion control method. The approach to design is identical to that for oil pipelines except that there is no pH buffering capacity in the condensed water in wet gas lines. This must be taken into account when performing the corrosion rate predictions. If the flow regime is stratified or wavy, there may be a concern that corrosion inhibitor deployed into the continuous phase at the bottom of the pipe does not get transported to the top of line location. The corrosion processes occurring at the two locations are different as the transportation

Corrosion Inhibitor Deployment in Wet Gas Pipelines

74

THE USE OF PREDICTIVE MODELS DURING DESIGN

of water to, and subsequent removal of corrosion products from the top of line location is limited by the quantities of condensing water. There is no continuous water phase at this location in stratified/wavy flow and water is only present via condensation on to the pipewall. Under these circumstances, the water quickly becomes saturated with corrosion products, effectively stifling further corrosion and this can be used to advantage in the design of wet gas pipelines. The term top of line/bottom of line (TOL/BOL) ratio is used to describe the rate at which the top of line corrodes relative to the bottom of line, with the bottom of line rate being calculated using a standard CO2 modelling approach. A TOL/BOL ratio of 0.1 is used by BP. This does not rely on inhibitor availability and can therefore be assumed to occur 100% of the time. The bottom of line location requires inhibition and the predicted rate estimated using the availability model. The higher of the two rates will determine the required corrosion allowance.
Corrosion Inhibitor and Glycol Deployment in Wet Gas Pipelines

Glycol (or methanol) is often used as the hydrate preventer on a recycled basis, although this traditional approach to hydrate control is increasingly being replaced by once through, low dose systems. However, recycled systems will remain valid for older systems or those operating well within the hydrate envelope where low dose chemicals are not applicable. The use of glycol is beneficial as it is a corrosion inhibitor, albeit a relatively poor one. If glycol is used without the addition of corrosion inhibitor, there will be some benefit from the glycol. This is hard to quantify but Shells work produced a glycol correction factor which is described on page 25. However, if glycol and inhibitor are both used there will be little additional benefit from the glycol and it should be ignored for design purposes. Only the inhibitor availability factor should be used. The use of a glycol (or methanol) recycling system offers the opportunity for an alternative form of corrosion control - pH moderation. This technique has been used by Elf since the 1970s and works by artificially raising the pH of the water in the pipeline to high values (circa 6.0). This limits or arrests CO2 corrosion and therefore the pipeline can be constructed with reduced corrosion allowance. The system is economical to operate as the pH moderator, typically bicarbonate or MDEA is carried in the glycol and remains through the glycol drying process. However, the technique should only be used along with corrosion inhibition as pH moderation is not entirely successful at preventing localised corrosion. In effect, pH moderation expands the application of carbon steel to more aggressive environments i.e. hotter and/or higher CO2 partial pressures.

75

THE USE OF PREDICTIVE MODELS DURING DESIGN

However, the technique has some drawbacks: 1. pH moderation relies on the existence of glycol recycling to provide the transport medium and to recycle the pH moderator. With the modern trend towards once-through, low dose hydrate inhibitors many new wet gas pipelines will not have glycol recycling facilities. Once-through dosing of pH moderator is unlikely to be economic as 2,500 ppm bicarbonate or 500 ppm MDEA may be required in the water phase to achieve the required pH shift. MDEA costs circa US$4 per kg and therefore treating condensed water would cost circa 30 cents per barrel on a once through basis. 2. If formation water is produced along with the gas then the artificially high pH will increase the scaling tendency of the water. This can have serious consequences and may require the termination of the pH moderation programme. Multiphase Flowlines Multiphase flowlines are the most arduous application for corrosion inhibitors. This mode of transporting fluids is set to increase further in BP with the development of long reach tie-backs to existing platforms and large developments on land such as Colombia and Algeria. Flowlines are an arduous application for corrosion inhibitors for two main reasons: 1. The fluids are unstabilised and therefore contain acid gases such as CO2 at high partial pressures, along with water. In contrast, export pipelines transport more benign fluids that have had the bulk of such corrodents removed. 2. The flow regimes in multiphase flowlines vary widely and the velocities and attendant liquid forces can reach high levels. This increases uninhibited corrosion rates and increases the concentration of inhibitor required to achieve acceptably low corrosion rates. Very low velocities are also a concern and the optimum mean velocity for such flowlines is believed to lie between 1 and 10 m/s. Below this velocity range, water drops out and deposits can accumulate at the 6 oclock position, preventing inhibitor reaching the pipewall - see page 46.

76

THE USE OF PREDICTIVE MODELS DURING DESIGN

Corrosion inhibitors are not effective under such circumstances, particularly if there are solids or bacteria in the line and therefore corrosion rates are increased. Corrosion rates can be controlled above 10 m/s, but at higher operating costs. Figure 24 gives a graphical representation of the affects of flow velocity on inhibited CO2 corrosion rates.
Figure 24: A Graphical Representation of the Effect of Flow Velocity on Inhibited CO2 corrosion.

Corrosion Risk

Effect of increasing [CI] High risk of water drop and under-deposit corrosion

0 0 5 10 Flow Velocity - m/s 15 20

Figure 24 is only a qualitative representation and velocity is not the only criterion controlling the flow element of CO2 corrosion. Mixture density is also important, with denser fluids giving rise to higher corrosion rates. Higher velocities can therefore be tolerated in systems with high GORs than in similar systems with low GORs. It is often convenient to design using Cfactors, defined in API RP 14E because the erosional velocity is often the limiting velocity for flowlines. Although C-factors specifically relate to erosion and not corrosion they usefully represent the forces acting on the pipewall and therefore the forces causing enhanced corrosion rates. For carbon steel, BP use a C factor = 135. MaximumFlow Velocity = C Mixture density

..where maximum flow velocity is in ft/s and mixture density is in lb/ft3.

77

THE USE OF PREDICTIVE MODELS DURING DESIGN

The relationship between flow and corrosion rate will be unique for each system and will be difficult to estimate at the design stage. However, the designer should accept that high velocities increase the risk of high corrosion rates and should design accordingly. The level and sophistication of corrosion control and monitoring systems must reflect the potential for corrosion to occur and this in turn will depend heavily of the flow regime. This should be handled using the approach developed for Inhibitor Availability, based on categories 1 to 5. The impact of flow velocities corresponding the C factors > 100 can be considered as an increase in risk and the category defined on the bais of predicted corrosion rates changed accordingly - see Table 14.
Table 14: The Impact of High Fluid Velocities on the Categorisation of Corrosion Risk

<100

C Factor 100 to 135

>135

Design Operation

No change Yes

+ 1 Category + 1 Category

No Possibly1

Note that operating at C factors > 135 should only be considered where there is sufficient operational experience in the asset to confidently state that erosion or corrosion are not occurring at unacceptable rates at C=135. C factors > 135 should not be used during design but may be considered as a de-bottlenecking measure if successful experience has been gained. 'Successful experience' is likely to require several years of operation with at least one intelligent pig inspection of the flowline after operating at close to C = 135.

78

THE USE OF PREDICTIVE MODELS DURING DESIGN

Intelligence Pigging Guideline Within BP, there are no fixed policies on the frequency of intelligence pig surveys (IPS) of pipelines and each individual case should be examined on merit. It is important to note that intelligence pigging surveys are just one element of a toolbox for the management of pipeline integrity. They are complementary to the full range of other pipeline integrity and monitoring techniques, for example, wall thickness checks, corrosion coupons and corrosion inhibitor injection monitoring. It is recommended that pipelines at risk of corrosion are designed to be piggable, with the requirement for permanent pig traps being determined according to the required frequency for operational pigging and intelligence pigging. For any pipeline, the need for, and frequency of inspection depends on a number of factors:

r the known or anticipated corrosion risk (which this document deals with); r the sensitivity of the inspection tools available to detect the anticipated defect types; r the corrosion allowance and whether the pipeline can be accessed for repairs; r the environmental risk; r local pipeline regulations; r the strategic importance of the pipeline and the associated political environment;
There are three main types of pipeline inspections which may be categorised as follows:

r A Baseline Survey; r an Early Inspection; and r a Routine Survey.


Baseline Survey

A Baseline Survey is carried out prior to pipeline commissioning, with the principal objective of detecting material defects and construction anomalies. Baseline surveys are primarily intended to detect dents, or wrinkles, and so geometry pigs are normally used (e.g. caliper device), these pigs are not normally considered to be intelligence pigs.

79

THE USE OF PREDICTIVE MODELS DURING DESIGN

BP would not generally recommend a baseline intelligence pig survey for pipelines. It is normally considered that the pipeline hydrotest is a sufficient demonstration of the pipelines initial fitness for service. A true baseline intelligence pig survey may be required for pipelines transporting highly sour fluids, where there is some doubt over the steels ability to resist HIC. In these circumstances it may be justified to inspect for initial laminations in the pipeline steel. Such laminations may grow or blister during operation, and so a baseline measurement of such features can prove useful in assessing the integrity of the pipeline in later life. BP has not found it necessary to carry out a baseline IPS in any of its pipelines. However, there is increasing pressure on the assets from regulators to carry out such inspections. A normally satisfactory compromise is the Early Inspection.
Early Inspection

An Early Inspection would be carried out 1 - 3 years after commissioning. The objective of this survey is to verify the absence of corrosion in a pipeline where a new corrosion prevention strategy is being implemented, or when the operating conditions are particularly severe. In the context of this document, pipelines in corrosion categories 4 and 5 would certainly warrant an early inspection. The case for a category 3 pipeline having an early inspection should also be considered. An early inspection is similar to a baseline inspection, but it is carried out after some operating life has been accumulated. The objective of an early inspection is to confirm that the corrosion management philosophy is operating satisfactorily before any significant damage occurs to the pipeline. For example, an early inspection of the Miller Gas System (sour with high CO2) was carried out after approximately 1 year of operation and this confirmed satisfactory corrosion management performance. The data from an early survey can be used later in the pipelines life to provide information on when the damage was initiated.

Routine Survey

A routine inspection is carried out to confirm the on-going integrity of a pipeline which has a known corrosion risk. Clearly, the frequency of this inspection will vary from pipeline to pipeline. This survey is used to monitor known defects or confirm the absence of significant corrosion. Pipelines in corrosion categories 1 to 5 should all be considered for routine surveys.

80

THE USE OF PREDICTIVE MODELS DURING DESIGN

When To Inspect a Pipeline

Where it is feasible, it is recommended that a caliper inspection is carried out on each new pipeline during the commissioning procedure. This will confirm the good workmanship of the pipeline and remove concern about dents in the pipeline. Dents can pass hydrotest, but fail by fatigue later during the life of the line. Baseline intelligence pig surveys are not generally recommended. To determine when to first inspect a new pipeline, it is necessary to consider the corrosion risk and the uncertainty in prediction of the corrosion rate. Three simple principles can be used to determine when to carry out the first pig inspection:

New Pipelines

r It should not be before a time when one would expect to detect some corrosion if the corrosion rate is in line with the pessimistic estimate; r It should be before our pessimistic estimate of when the first failure may occur; r It should be before we expect widespread corrosion to occur, which would result in major repair programme.
To quantify this timing, the pipeline operator can apply a number of methods of increasing sophistication, reliability calculations are suited to this type of assessment. However, as a first pass the following simple method is recommended: Make an estimate of the most likely and the pessimistic corrosion rates. These should be based on the corrosion model described here, taking in to account the influence of corrosion inhibitors (if applicable) and the likely effectiveness of the inhibitors. The probabilistic approach to corrosion monitoring can be helpful here, taking the P50 and P90 (or P10) corrosion rates as the most likely and pessimistic rates. The first inspection should not be before. Calculate the time taken for the pessimistic corrosion rate to reach 1mm in depth or the detection threshold of the inspection tool being used (typically 10% of nominal wall thickness). Use the greater time period as the earliest recommended inspection time.

81

THE USE OF PREDICTIVE MODELS DURING DESIGN

The first inspection should not be after. Calculate the retiral thickness of the pipeline according to the ruling pipeline design code - the code thickness. Calculate the tolerance of the pipeline to long corrosion defects. For pipelines operating at 72% SMYS, the BP Guidelines for the assessment of corroded pipe allow a further 20% loss of the Code thickness (BP's "Transmission Pipelines to BS8010", 1st June 1992). The BP 1st Pass thickness is normally calculated as 0.8 * code thickness. Calculate:

r the time that the pessimistic corrosion rate will reach the BP 1st Pass Thickness, r half the time that the best guess corrosion rate will reach the BP 1st Pass Thickness.
The earliest of these dates is the latest intelligence pig inspection date. The procedure is shown pictorially in Figures 25 and 26.
Figure 25: A Pictorial Representation of how to Determine the Timing of a Pig Inspection Run.
Corrosion Levels Wall Thicknesses
Nominal Detection Level Design Allowance Actual Allowance Failure Point

SAFE
Code

SAFE UNSAFE
pessimistic estimate best estimate

BP 1st Pass long Rupture

FAILURE

0
earliest latest inspection inspection date date (1)

time
widespread corrosion expected (divide time by two)

82

THE USE OF PREDICTIVE MODELS DURING DESIGN

Figure 26: A First Pass Method for Determining the Timing of the First Intelligence Pig Inspection.

Calculate the most likely and worst case corrosion rates (P50 and P90 rates if using the probabilistic corrosion model)

Determine when the most likely rate leads to wall losses > 10%

Determine the retiral wall thickness using BPs 1st pass method for long defects Calculate when the worst case rate means the retiral limit will be reached

5 2
Determine when the most likely rate leads to wall losses of 1 mm

Calculate when the most likely rate means half the retiral limit will be reached

Take the greater of the times calculated in Steps 1 and 2 This is the earliest date for inspection

Take the lesser of the times calculated in Steps 5 and 6 This is the latest date for inspection

The actual inspection date chosen should fall between these limits. The final selection of date will depend the factors outlined above i.e.

r the known or anticipated corrosion risk; r the sensitivity of the inspection tools available to detect the anticipated defect types; r the corrosion allowance and whether the pipeline can be accessed for repairs; r the environmental risk; r local pipeline regulations; r the strategic importance of the pipeline and the associated political environment;

83

THE USE OF PREDICTIVE MODELS DURING DESIGN

Repeat Inspection Intervals

A similar method is used to determine the minimum inspection interval for pipelines without a severe corrosion problem. The relatively low accuracy of even the high resolution intelligence pigs compared with other NDT techniques, means that pigs are not well suited to the measurement of corrosion rate. Statistical techniques have been applied to pig inspection results. However these will result in a high degree of uncertainty in measured corrosion rate unless there is a reasonable period of time between inspections. For example, the time to the next intelligence pig inspection survey could be determined as follows. If the defects identified in the early survey are indeed corrosion defects, then one should carry out the next inspection when the predicted growth exceeds the tools ability to confidently measure differences in wall thickness. If an inspection tool has an accuracy of 10% of pipewall thickness then the inspection should be carried out when the estimated total loss in wall thickness due to corrosion has exceeded: 2 x 10% i.e. 14.1%, which is equivalent to 1.8 mm on a 12.7 mm thick line. The reason for this is that the error in the measurement of corrosion (a differences in wall thickness) is approximately 2 times the error in each wall thickness measurement. For pipelines with significant corrosion, the timing of the next inspection depends on when it is anticipated that the corrosion depth will reach a "retiral" limit. For onshore pipelines, the owner has the opportunity to carry out local inspections and repairs at relatively low cost. In this instance, an inspection programme can be put in place to monitor a number of the severest defects in order to judge when repair / replacement / derating is necessary. This point monitoring may be used to reduce the required frequency of IPS. For offshore pipelines with significant corrosion, where inspection and repair is costly, there will be a tendency to carry out IPS more frequently than outlined above. It should be understood that inspections carried out more frequently than the minimum recommended frequency may not be able to generate reliable corrosion rate data. In these instances only with careful consideration, should forecasts of pipeline integrity be made from pit depth changes from inspection to inspection. In order to avoid over-pessimism in forecasts it is important to consider other sources of information on possible corrosion rates (e.g. corrosion model predictions / experiments; topsides inspection results).

84

THE USE OF PREDICTIVE MODELS DURING DESIGN

Typical Inspection Intervals

For oil transmission pipelines, where fluid corrosivity is being monitored on a routine basis, the pipeline is in a reasonable condition and thought to be at low risk, a frequency of once every 5 years would be typical. Examples of this are the new Forties MOL and the existing Ninian MOL, which are both subsea lines in the North Sea. When a good corrosion management track record has been established, assets are tending to increase this interval. For significantly corroded pipelines, where the pipeline is nearing the end of its life, inspections may be carried out as often as annually. For dry gas pipeline systems that are tightly controlled, inspections would be carried out after indications of potential problems from other sources: topsides corrosion, failure to meet dew point spec, water carry over into the pipeline etc. For example, BP has operated a dry gas pipeline (Gyda field in Norway) since 1986, without yet requiring an intelligence pig inspection, because of the low risk of internal corrosion in this pipeline. For more information on this topic contact Will McDonald ( Sunbury x4014 ) or Jim Corbally ( Sunbury x2774 ) of the SPR Transportation Team.

85

THE USE OF PREDICTIVE MODELS DURING DESIGN

Applying Models to Process Equipment


Crude Oil Stabilisation Trains

As stabilisation trains take fluids from the flowlines, they will naturally benefit from the injection of any corrosion inhibitors upstream to protect the flowlines. However, there are locations within the stabilisation units where inhibitors will not work well and alternative means of corrosion control should be employed. Inhibitors rarely work well under low velocity or stagnant conditions, such as at the base of separators, tanks or in instrument bridles. Deposits can form in such locations preventing inhibitors getting to the metal surface. This becomes relevant at velocities below 1 m/s and either internal coatings and anodes (vessels, tanks) or stainless steel piping (instrument bridles) should be used. Carbon steel is suitable for drain lines, downstream of an isolation valve. Gas compression systems fall into two categories; wet gas compression and dry gas compression. Some systems are wholly wet gas, such as Pedernales, Venezuela and the Long Term Test facility at Cusiana, Colombia. The majority of systems are wet up to an intermediate stage of compression at which point the gas is dried, normally in glycol contactors at approximately 500 psi. Once the gas is dried, corrosion is not a major concern and a minimal corrosion allowance is normally specified to account for periods when gas dryers operate off-specification or for external corrosion. In wet systems, corrosion will occur whenever the gas falls below its water dew-point. This can be predicted using flowsheet simulation packages such as Genesis but there are some general guidelines which make the task more straight forward.

Gas Compression Systems

Pipework Downstream of Compressors

The gas entering a compressor will have come from either a vessel or knock out pot. The gas will therefore be in equilibrium with water and hydrocarbon liquids and there should be zero or negligible liquids present. The action of compressing the gas will heat it, raising it above the dew-point and thereby removing any traces of liquid water. The pipework downstream of compressors is therefore not at risk from internal corrosion and a moderate corrosion allowance (1 - 2mm) will suffice to account for external corrosion. The exception is small bore instrument tappings where the gas may cool to below its dew-point, causing corrosion. Greater corrosion allowances or stainless steels should be used in these locations.

86

THE USE OF PREDICTIVE MODELS DURING DESIGN

Wet Gas Coolers

Gas is typically cooled between each stage of compression. Downstream of compressors, liquid water will not re-appear until the gas is cooled to below its dew-point. This will occur some way into the cooler. If the cooler has carbon steel tubes it is worth calculating the temperature at which this will occur as the site of water condensation can be the location of worst case corrosion and will therefore determine the life of the coolers. As the following example, in Table 15 from one BPX asset shows, the dew-point temperature can be closer to the gas exit temperature than the entry temperature. If the entry temperature had been used for the corrosion rate predictions, they would have been unnecessarily conservative.
1st Stage Discharge Cooler 2nd Stage Discharge Cooler 3rd Stage Discharge C ooler 4th Stage Discharge Cooler

Table 15: Operating Conditions and Corresponding Corrosion Rates for Discharge Coolers
Temperature - in Temperature - out Pressure CO2 content PCO2 pH range Water dew point Water content Tube size Wall thickness Pred. Corr. Rate Inhib. Corr. Rate

243oF 110oF 220 psig 0.2 mole% 0.44 psia 4.65 - 4.95 147oF 737 lb./MMscf 1" x 16g 1.5 mm 0.5 mm/yr. 0.05 mm/yr.

240oF 110oF 567 psig 0.2 mole% 1.13 psia 4.47 to 4.75

222oF 110oF 1334 psig 0.2 mole% 2.68 psia 4.32 to 4.55

226oF 110oF 3310 psig 0.2 mole% 6.63 psia 4.21 to 4.38 132oF 64 lb./MMscf 5/8" x 16g 2.75 mm 1.6 mm/yr. 0.2 mm/yr.

144oF 136oF 299 lb./MMscf 124 lb./MMscf 5/8" x 16g 1.5 mm 0.9 mm/yr. 0.09 mm/yr. 3/4" x 16g 1.5 mm 1.16 mm/yr. 0.1 mm/yr.

87

THE USE OF PREDICTIVE MODELS DURING DESIGN

Carbon steel is rarely a good choice for the tubes of coolers in wet gas service for the following reasons: 1. The equipment is critical as it handles flammable gas at high pressure within the facilities. 2. The thermal requirements of the cooler exclude the use of significant corrosion allowances and therefore coolers typically have thin walled tubes. 3. The high gas velocities and highly turbulent flow regimes mean corrosion inhibitors are unlikely to work well. 4. Inhibitor may need injecting downstream of each compressor as it may be lost with the liquids at each knock out pot, making inhibitors uneconomic. 5. On-line inspection of the tubes of airfin coolers is difficult. More suitable materials for the tubes include 316L, duplex or super duplex stainless steels. If necessary, carbon steel can be used for the tube sheets to reduce costs with a suitable corrosion allowance incorporated.
Glycol Contactors

Glycol contactors are an example of equipment that, on the face of it, may suffer excessive internal corrosion due to the combination of gas below its dew-point, high pressures and carbon steel construction. However, operating experience has shown this to not be the case as the large volumes of glycol effectively absorb the water and inhibit corrosion. Carbon steel is therefore a satisfactory material of construction although many projects go to the expense of internal coatings, such as epoxy phenolics, particularly for the lower sections. Although corrosion inhibitors are supplied to control corrosion in glycol contactors, their benefit is not quantified or proven. However, the control of the pH of the water/glycol moisture is important and chemicals (neutralisers) are available for this. During operation the pH of the fluid is reduced by the build up of organic acids. These result from the degradation and hydrolysis of the glycol during the heated, regeneration stage.

88

THE USE OF PREDICTIVE MODELS DURING DESIGN

Flow Velocities in Process Pipework The 1995 de Waard & Milliams corrosion rate prediction model relies heavily on the use of flow velocities to predict corrosion rates. If this model is used, guidance is required on typical velocities. The design guidelines used for the Project, or actual throughput rates and internal pipe sizes are the best sources of such information. If this information is not available then the following information can be used as it details typical limiting velocities used during the design of process pipework. This information is taken from two recent design guidelines used by Process Engineers for sizing of process pipework. They deal with maximum velocities and can therefore be used as worst case. Pipe sizes are based on several criteria, including the requirements to avoid vibration, deposition of solids, excessive pressure drop and erosion.
Flow Velocities in Single Phase Liquid Lines

Single phase liquid lines refers to pipework where system pressure is forcing liquid from higher pressure vessels to lower pressure vessels, drains or tankage. It does not refer to the suction or discharge of pumps. Maximum velocity = 5.0 m/s, with excursions up to 9 m/s. Flow should not exceed 5.0 m/s and should not be less than 1 m/s. The lower limit is to avoid deposition of solids. More detailed guidelines are summarised below. They are only to be applied to clean fluids - allowable velocities shall be reduced if solids are present.

Table 16: Maximum Velocities in Single Phase Liquid Lines

Nominal line size Process liquid general Hydrocarbon headers Hydrocarbon branches Water & water solutions Liquid to reboiler Side stream drawoff Gravity flow Refrigerant lines

Maximum Velocity (m/s) 5 5 5 3.5 1.25 1.25 1 0.6 9 9 9 9 3 2.5 1.25

89

THE USE OF PREDICTIVE MODELS DURING DESIGN

Flow Velocities in Pumped Liquid Lines

The flow velocity in pumped liquid lines is strongly dependent on pump type and line size. Centrifugal pumps and large line sizes can handle higher liquid velocities than reciprocating pumps and small line sizes. If the pump type is unknown, it is safer to assume a centrifugal pump for the purposes of corrosion rate calculations. If the line size is not known, the following velocity range can be used. If the line size is known, Tables 17 and 18 give more information. Centrifugal Pumps Suction 1 to 2.4 m/s Discharge 1.8 to 5.5 m/s, excursions up to 9 m/s. Reciprocating Pumps Suction 0.3 to 0.6 m/s Discharge 1 to 1.8 m/s

Table 17: Maximum Velocities in Lines to and from Centrifugal Pumps

Service

Max. Velocity m/s Normal

Max. Velocity m/s Limit

up to 3" 4" 6" 8" 10" 12"

Suction 1 1.4 1.5 1.8 2.1 2.4

Discharge 1.8 2.4 3 4.3 4.9 5.5

Table 18: Maximum Velocities in Lines to and from Reciprocating Pumps

Speed RPM

Maximum Velocity (m/s)

up to 250 251-330 over 330

Suction 0.6 0.5 0.3

Discharge 1.8 1.4 1

90

THE USE OF PREDICTIVE MODELS DURING DESIGN

Flow Velocities in Multiphase Lines

In multiphase lines, use the limiting velocity defined by API RP 14E. BP use a C factor of 135 for carbon steel - see p77. Velocities should not exceed 75% of the 'critical flow velocity'. Critical flow in multiphase systems is analogous to sonic flow in single phase systems. A general limit of 18 m/s is applied to gas piping to avoid pipe vibrations. Compressor surge/recycle lines, relief valve inlets etc may operate at substantially higher velocities - see Table 19. However, pipework to and from reciprocating compressors typically has a lower velocity limit of 12 m/s. Service
< 15 psia (vacuum) 0 - 100 psig 100 to 500 psig 500 to 2000 psig Flare

Flow Velocities in Vapour or Gas Lines

Table 19: Vapour Line Sizing Criteria

Velocity m/s
61 to 152 46 to 61 30 to 46 30 to 38 0.5 to 0.8 Mach

In general, vapour piping is sized in terms of pressure drop, rather than maximum velocities.

91

THE USE OF PREDICTIVE MODELS DURING DESIGN

Economic Tools To Use During Materials Selection Corrosion modelling will give a good indication of the probability of failure of equipment in service due to internal conditions but will not help in determining the economic consequences of such a failure or the operating costs involved with avoiding or managing such a failure. Even if corrosion models predict short times to failure, it may be economic to plan for replacement or repair of carbon steel equipment late in field life rather than to invest in a more robust solution on Day 1. Alternatively, inhibitors may be a technically feasible solution for process pipework but economically and logistically, protecting large numbers of short lengths of pipework may be impractical and corrosion resistant materials may be a better choice. The technique of life cycle costing (LCC, also known as whole life costing) helps in this assessment by converting future costs into current monetary value and thereby allowing direct comparisons with capital costs. To carry out accurate, meaningful and useful LCCs the Materials or Project Engineer must have: 1. An understanding of the economic factors driving the decision, such as discount rates, rates of return on investment and net present values. 2. The design life and production profile of the development. 3. An assessment of future costs based on similar developments over several years. 4. An understanding of the important economic drivers for the Project, such as the balance between capital and operating costs. This in turn will be determined by the economic terms under which the licence was awarded. Gathering the necessary data for accurate LCCs is a major task and a guideline document is available [17]. In some cases, the cost of materials are relatively minor and the costs of installation far outweigh them. Expensive sub-sea wells are an obvious example of where workovers are to be avoided due to a materials failure. In such cases it is common to select robust materials in order to protect against a repeat of the high installation costs but there are many examples where the answer is less clear cut. The key question is, when is investment in corrosion resistant materials justified?

92

THE USE OF PREDICTIVE MODELS DURING DESIGN

Corrosion models clearly have an input to this but can not provide the complete answer. Corrosion models are normally used as a materials selection tool and taking an extreme example, if there were no consequences of a failure there would be no justification in investing in corrosion resistant materials. An investment in corrosion resistant alloys (CRAs) aims to protect against the consequences of a failure and therefore materials selection must consider the consequences in the decision making process. Consequences may include economic, health, safety or environmental impacts or all four but in most cases all consequences can be related to a financial impact. Example: A flowline is to transport corrosive fluids from a remote well-site to the processing facilities. The route includes a major river which provides local communities with water for consumption and agriculture. The river crossing requires directional drilling and is therefore expensive. The material selected for the majority of the flowline is carbon steel with a suitable corrosion allowance but it is recognised that localised failures and repairs may be required late in field life. What material should be used for the river crossing? The decision can not be based solely on the corrosivity of the fluids as the consequences of a failure under the river crossing is clearly far greater than a similar failure on land. A method of evaluating the consequences of such a failure is required and from this a method for determining how much it is worth investing on Day 1 to prevent a failure several years later. The Expected Value technique does this and is covered in detail in ref 17. The technique quantifies what has been done subjectively for many years: materials selection becomes more conservative as the consequences of a failure increase. This is the main reason corrosion resistant materials are used more extensively downhole and sub-sea than on land - it is not the fluids that are significantly different but the economic drivers.
Table 20: Categories of Equipment, Classified by the Proportion of Materials Cost to Total Installation Cost

Location/Equipment Type

Materials Cost as % of Whole < 3% ~ 10% ~25% ~ 30% > 30%

Material Selection

Subsea wells Land wells / sub-sea flowlines Flowline road / river crossings Buried land lines Surface running land lines

Most conservative

Least conservative

93

THE USE OF PREDICTIVE MODELS DURING DESIGN

The expected value technique is represented graphically in Figure 27. It assesses the economic costs and benefits of two or more choices. In assessing each option, the technique allows for the possibility of failure and a probability is assigned to each outcome (failure or no failure). Probabilities of failure will be higher for a carbon steel system than for an equivalent CRA system and corrosion modelling helps to determine this. The costs of each outcome consist of :

r Capital costs (no failure case) r A combination of capital and operating costs r The above plus repair or replacement costs
For fair comparison, the costs are converted to present day values (NPVs). The costs associated with each outcome are multiplied by their probability to produce the estimated value.
Figure 27: Expected Value Technique
No Failure 99% Install CRA river crossing Failure 1% NPV Cost = $1.0+$0.48+$23.75 = $25.23 Choose lowest EV i.e. CRA river crossing NPV Cost = $1.0 EV for CRA $1.31 million
(0.99 x 1) + (0.01 x 25.23)

Which river crossing material ?

No Failure 80% Install C-Steel river crossing 20% Failure NPV Cost = $0.6+$0.29+$23.75 = $24.64 NPV Cost = $0.6 EV for C-steel $5.85 million
(0.8 x 0.6) + (0.2 x 24.64)

94

References

1.

C de Waard, U Lotz, D E Milliams, "Predictive Model for CO2 Corrosion Engineering in Wet Natural Gas Pipelines", Corrosion, 47 (1991) 976 C de Waard, U Lotz, "Prediction of CO2 Corrosion of Carbon Steel", NACE Corrosion 93, New Orleans, paper 69 C de Waard, U Lotz, A Dugstad, "Influence of Liquid Flow Velocity on CO2 Corrosion : A Semi-Empirical Model", NACE Corrosion 95, Orlando, paper 128 C de Waard, D E Milliams, "Carbonic Acid Corrosion of Steel", Corrosion, 31 (1975) 177 R H Newton, "Activity Co-efficients of Gases", Industrial and Engineering Chemistry, March 1935, 302-306 L W Jones, "Corrosion and Water Technology", OGCI Publications, Tulsa, USA, 1988, p14-15 J G Stark, H G Wallace, "Chemistry Data Book", J Murray Ltd, London, 1978, p 60-61 I R McCracken, C G Osborne, D Harrop, "Carbon Dioxide and Corrosion in Forties", Sunbury Report No PEB/122/89, dated December 1989 J E Oddo, M B Tomson, "Simplified Calculation of CaCO3 Saturation at High Temperatures and Pressures in Brine Solutions", J of Petroleum Technology, 34 (1982) 1583 L A Rogers, M B Tomson, "Saturation Index Predicts Brine's ScaleForming Tendency", Oil and Gas Journal, April 1 1985, p 97 R G Chapman, "pH Models for Corrosion Rate Predictions", Sunbury Report No POB/025/96, dated June 1996 M J J Simon Thomas, P B Herbert, "CO2 Corrosion in Gas Production Wells: Correlation of Prediction and Field Experience", NACE Corrosion 95, Orlando, paper 121

2.

3.

4.

5.

6.

7.

8.

9.

10.

11.

12.

95

REFERENCES

13.

J J Carroll, J D Slupsky, A E Mather, "The Solubility of Carbon Dioxide in Water at Low Pressure", J Phys Chem Ref Data, 20 (1991) 1201 - 1209 A J Ellis, R M Golding, "The Solubility of Carbon Dioxide above 100C in Water and in Sodium Chloride Solutions", Amer J of Sci., 261 (1963) 47-60 S Takenouchi, G C Kennedy, "The Solubility of Carbon Dioxide in NaCl Solutions at High Temperatures and Pressures", Amer J of Sci, 263 (1965) 445 - 454 S D Malinin, "Thermodynamics of the H2O - CO2 System", Geochemistry International, 10 (1974) 1060 - 1085 D M E Paisley, "Selecting Materials for Wealth Creation: A Materials Selection Philosophy based on Life Cycle Costs", BP Sunbury Report No. ESR.97.ER.005, 10th Jnauary 1997 D Vedapuri "Studies on Oil-Water Flow in Inclined Systems" April 1997 Progress Report, Section 9. Ohio University Multiphase Flow and Corrosion Project. A J McMahon and S Groves, "Corrosion Inhibitor Guidelines: A Practical Guide to the Selection and Deployment of Corrosion Inhibitors in Oil and Gas Production Facilities", BP Sunbury Report No. ESR.95.ER.050, April 1995

14.

15.

16.

17.

18.

19.

96

Installation of the Cassandra 98 Excel Workbook


Description The Cassandra 98 work book was written in Microsoft Excel for Windows 95, version 7.0a. It may not run in earlier versions of Excel. Automatic Installation If for any reason this does not succeed, try the Manual Installation procedure described below. 1. 2. 3. 4. 5. 6. Insert the disc into the disc drive Click on the Start button Click on the Run option Using the Browse feature select A:\Install.Exe In the Run window click on OK Follow the Instructions.

Once complete the work book should be opened using the following sequence: 1. 2. 3. 4. Start Programs Cassandra Cassandra 98

The first time the work book is used the message This document contains Links will appear. Click No to this. Continue at step 5 in the Manual Installation procedure described below. If not already present, the automatic installation will create the following folders with files in them: 1. C:\Xlph 2. C:\Data\Cassandra 3. C:\Windows\Start Menu\Programs\Cassandra In addition it will place the file Xlph.ini in the root directory ( c:\ )

97

INSTALLATION OF THE CASSANDRA 98 EXCEL WORKBOOK

Manual Installation In the root directory of the disc there is a folder called Files. This folder contains two files ( Cassandra 98.xls and Xlph.ini ) and a folder ( Xlph ) in the root directory. The Xlph folder contains seven folders: 1. 2. 3. 4. 5. 6. 7. Phreeqe.dat Readme.doc Xlph.inf Xlph.out Xlph.xla xlph.xla Xlphdemo.xls

It is suggested that these instructions are visible during loading so that they can be referred to easily during the loading process. The instructions must be followed precisely to ensure that the installation is successful. It is suggested that Windows Explorer or File Manager be used for sections 1 to 3 below.
Installation

1. Copy the Xlph folder into the root directory of the C: drive. This should give the following structure: 1. 2. 3. 4. 5. 6. 7. C:\Xlph\Phreeqe.dat C:\Xlph\Readme.doc C:\Xlph\Xlph.inf C:\Xlph\Xlph.out C:\Xlph\Xlph.xla C:\Xlph\xlph.xla C:\Xlph\Xlphdemo.xls

2. Copy the Xlph.ini file into the root directory of the C: drive to give C:\Xlph.ini 3. Copy the Cassandra 98 file to your preferred location such as the Desktop, the root directory or another folder. For example: C:\Cassandra 98.xls 4. Start Excel 5. On the Menu bar click Tools. 6. On the drop down menu click Add-Ins. 7. In the Add-Ins box click on Browse. 8. In the Look in box select (C:). 9. Select the xlph folder and click Open. 10. Select the Xlph.xla file and click OK. 11. In the Add-Ins box click on OK. 12. If not already open, Open the Cassandra 98.xls file.

98

You might also like