SPE 147244 A Flow Assurance Study On Elemental Sulfur Deposition in Sour Gas Wells

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

SPE 147244

A Flow Assurance Study on Elemental Sulfur Deposition in Sour Gas Wells


Yula Tang, Joe Voelker, Chevron Asia South E&P; Cengizhan Keskin, Zhenggang Xu, Bin Hu, SPT Group;
Changqing Jia, China National Petroleum Company

Copyright 2011, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Annual Technical Conference and Exhibition held in Denver, Colorado, USA, 30 October–2 November 2011.

This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents of the paper have not been
reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect any position of the Society of Petroleum Engineers, its
officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written consent of the Society of Petroleum Engineers is prohibited. Permission to
reproduce in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract
This paper summarizes a flow assurance study for elemental sulfur deposition in the tubing for the Chuandongbei Gas Project
(CDB), a greenfield sour dry gas development project in Sichuan, China. The project is a co-venture of Chevron and China
National Petroleum Company (CNPC). The development contains several fields, all containing dry gas with 8-17% H2S, and
8-10% CO2.
Elemental Sulfur (S8) dissolved in the gas may precipitate in both the near-well reservoir region, and tubing, within the
pressure and temperature range predicted for gas well flow in the fields. The precipitate may form gas flow restrictions. An
offset operator producing gas from the same reservoir interval and having similar composition has reported flow problems
attributed to S8. This study focused on the prediction of S8 deposition in the tubing, during both production and shut-in
periods.
Numerical transient well pressure and temperature modeling with the OLGA wax deposition module was used to predict
precipitation and deposition of S8 in the tubing. Presently, no other S8 deposition model is available. The model estimates the
pressure and temperature gradient between the bulk fluid and tubing wall, and molecular diffusion rates through the laminar
sub-layer of the fluid velocity profile. An S8 phase equilibruium model calibrated by measured phase behavior from a
laboratory synthetic gas having the composition of CDB field measured surface gases was used to generate the S8 phase
diagram.
The study indicates that shut-in periods present the greatest S8 deposition risk: suspended sulfur precipitate accumulates at
bottomhole during shut-in periods, possibly forming a flow restriction. Mitigation through solvent treatment applied after
shut-in periods is therefore planned. S8 deposition on tubing walls during flow periods was found to present relatively low
flow assurance risk.
This study also provides operational design recommendations for well start up, sustained flow, and shut-in periods, as
well as flow assurance mitigation design.

Introduction
Elemental sulfur occurs in solid, liquid and gas phases, depending on pressure, temperature, and gas composition. Solid
sulfur exists in rhombic crystalline form when the temperature is less than 96 ºC, in monoclinic crystalline form up to 118 ºC,
and in gas form above 120 ºC. Elemental sulfur in gas or liquid phases will have a ring structure with eight atoms linked
together (S8), that will break down above 157 ºC. As most reservoirs do not exceed this temperature, elemental sulfur is
commonly referred to as S8. Commonly the solid phase is also referred to as S8, as is the case in this paper.
S8 occurs as a dissolved species in virtually all deep sour gas reservoirs. Studies indicate the solubility of S8 is positively
correlated to H2S concentration.4,5,7,10 Sulfur precipitation during gas production is induced by a reduction in S8 gas phase
solubility, below the in-situ concentration, as a result of decreases in pressure and temperature.
Solid sulfur deposition can occur in the reservoir, tubing, and surface facilities. Prediction and mitigation of sulfur
deposition are crucial to the economic viability of sour gas fields. Sulfur deposition risk occurs in the tubing during the
process of well start-up, stable production, and shut-in. It is critical to evaluate the risk and seek mitigation at the planning
stages of field development.
Flow assurance risk is assumed to develop from two possible S8 precipitation processes in the tubing, in this study:
• Continual aggregation of S8 on the tubing wall during start-up and production: during well start-up or stable production,
the tubing wall is cooler than the center of hot gas stream. S8 will tend to precipitate from the boundary layer that
2 SPE 147244

adheres to the cooler tubing wall. As the boundary layer has the lowest velocity in the stream, S8 precipitate may
deposit on the tubing wall. Growth of the deposition thickness may result in a flow restriction.
• S8 precipitation during well shut-in: precipitate resulting from wellbore cool-down may adhere to the tubing wall, or fall
from suspension and accumulate at the wellbore bottom. Both types of accumulation may become flow restrictions.
Mitigation of either process is assumed to be:
• Producing the wells at a rate sufficient to exceed the critical aggregate growth velocity,
• Solvent treatment soaking during production, startup and shut-in.
To accurately predict the sulfur deposition by these two processes, three models were used:
1) A phase equilibrium model
2) A flow/heat dynamic model to estimate transient pressure and temperature in the wellbore,
3) A precipitation/deposition model to estimate sulfur deposition rate on the inner surface of the tubing and accumulated
in the wellbore bottom.

CDB and Offset Field Sulfur Occurance


The productive sour gas reservoir at CDB is the Trassic Feixianguan (T1f). The T1f achieved its deepest burial in the
Cretaceous, with a temperature exceeding l50ºC. This temperature enabled the Thermochemical Sulfate Reduction (TSR)
process, generating significant H2S concentrations, and solid phase elemental sulfur. The Himalaya uplift resulted in
shallowing and lower temperature, terminating TSR.
Sulfur deposits have been found in CDB T1f cores, within anhydrite nodules, calcite cements, and horizontal or low-angle
fractures, as shown in Fig. 1 and Fig. 2. Sulfur deposition is seen in older fractures, and less so in most recently formed vugs
and fractures.

Fig. 1. Sulfur deposition (yellow) in T1f core Fig. 2. Sulfur deposition in T1f core (a: anhydrite, c: calcite, d: dolomite, s: sulfur)

The potential volume of elemental sulfur precipitation upon production is highly uncertain at CDB. All gas samples
acquired to date are from surface equipment only, and therefore current in-situ estimates of S8 saturation are not useful.
Therefore, a conservative approach is used in this study – CDB sour gas is assumed to be saturated with S8 at reservoir
conditions.
Producing sour gas fields offset the CDB block, approximately 100 km southeast. The gases in these fields have similar
compositions to that in CDB.
Two of three sour gas wells in these fields have experienced sulfur deposition in surface productin equipment and in the
tubing. The two wells produce from the T1f reservoir. The outlet spool of the separator was the major point of S8 deposition,
as shown in Fig. 3. Sulfur deposits occurred in the initial stage of production ramp up. The third well produces from the
deeper Changxing formation and does not have sulfur deposition downhole or in surface equipment. Table 1summarizes
basic information for the three offset sour gas wells, in comparison with an existing typical CDB well, Well A. Fig. 4 plots
the S8 solubility curves based on the study by Alberta Sulphur Research Ltd. (ASRL),11 and the possible P/T pathes of the
wells production from reservoir to the surface conditions. Sulfur precipitation is induced by a reduction in the solubility of
the sulfur in the gas phase beyond its thermodynamic saturation point as a result of decreases in pressure and temperature.
Note that Well A-15 has extreamly high reservoir pressure thus it may have the highest S8 content (about 2.0 g/m3) in
reservoir conditions (9357 psia / 192F). The reported S8 depostion in the separator exit may be due to the combination of
excessive S8 content in the gas stream, and the low P/T at the surface when flowing through the separator restriction. Its
initial shut-in wellhead pressure SIWHP=7500 psia, and FWHP=7100 psia at 5.3 MMscf/d decreased quickly to
FWHP=1276 psig at 7 MMscf/d in less than three years (Oct., 2008). It is not clear if that was due to S8 reservoir/wellbore
plugging or due to reservoir depletion, as no detailed reservoir and geological information collected.
Well B-51 may not have high S8 content at the reservoir conditions (5365 psia and 181 F). However, the flow rate is
believed to be close to the mimumum hydrate formation rate based on composition calculation. Thus S8 depostion was found
SPE 147244 3

to be mixed with the hydrate plug in this well. Once hydrate plugging has formed, S8 will have more chance to deposite out
due to local P/T reduction (J-T cooling) through restriction (WHP=1830 psia, WHT=80 F).
Comparibly, offset Well C-X3 with higher rate than the other two wells, has no hydrate formation issue, neither apparent
S8 plugging in the well. It may have sufficient flow velocity and temperature to avoid S8 deposition in the well and in the
separator. However, it may have the potensicial to form S8 deposition in solid phase in the near wellbore reservoir (initial
reservoir P/T on the right hand side of the dashed line, Fig. 4) and it can damage the well deliverability.

Fig. 3 Elemental sulfur plug was found in the separator outlet spool in A-15 well

Sulfer Solubility  Phase Diagram  for CDB Field A, Predicted by ASRL

10000
Offset Field Well A‐15 
Reservoir P/T
9000 Well‐A, CDB Field A, 50 MMSCFD

Well‐A, CDB Field A, 10 MMSCFD

8000

Offset Field Well A‐15  1.0 g/m3 Offset Field 


inital WHP/WHT Well C‐X3 
7000
0.5 g/m3 CDB Field A  Reservoir P/T
Reservoir P/T

6000
0.1 g/m3
Pressure, psia

5000
Offset Field 
Well B‐51 
4000 WH Choke  Reservoir P/T
Upstream 

3000

WH Choke  0.01 g/m3
2000 Downstream
On the left, 
0.001 g/m3 S8 in liquid state;
On the  right, 
1000 S8 in solid state.
Offset Field Well A‐15   0.0001 
later WHP/WHT g/m3
0
0 20 40 60 80 100 120 140 160 180 200 220 240 260 280 300

Temperature,  F

Fig. 4 Sulfur solubility for CDB Field A gas, predicted by ASRL

Table 1: Offset high-H2S wells basic information


4 SPE 147244

Well A in Field A,
Well Name A-15 B-51 C-X3
CDB

Producing Formation Feixianguan (T1f) Feixianguan (T1f) Changxing (P2ch) Feixianguan (T1f)

TVD, m 4075 3670 5088 3500


Well Type Vertical Vertical Inclined Horizontal (535 m)
Completion Perf Perf Perf Perf
3.5" x 1348m + 2.875" x 3236.8 +
Tubing String 3.5" 4.5"
2.875" x 3894m 2.38" x 3630m
Sample location Surface Surface Surface Surface
Initial production date Oct., 2005 Apr., 2004 May, 2004 (test )
Initial gas rate, MMscf/d 5.3 2.8 10.6 107 (tested rate)
Initial reservooir pressure, psia 9357 5365 7250 5840
Initial reservooir temperature, F 192 181 230 195
03/07/2006,
Plugging date - No -
7/12/2006
High pressure
Mitigation method wash, glycol Bullheaded glycol No -
bullheading.
yellowish mixture
S8 plug observation Solid sulfur plug No -
with hydrate
tubing, wellhead
Separator gas choke, separator
Sulfur Deposition Location No -
outlet inlet, orifice plate
on gas meter
Hydrate +
Plugging material Element sulfur no no production yet
elemental sulfur
Compositions no production yet
C1 90.75% 89.85% 78.85% 82.28%
C2 0.08% 0.21% 0.05% 0.05%
C3 0.00% 0.02% 0.00% 0.01%
C4+ 0.00% 0.002% 0.000% 0.000%
H2S 7.0% 6.9% 8.9% 9.5%
CO2 1.6% 2.6% 11.9% 7.5%
N2 0.52% 0.39% 0.28% 0.60%
He 0.02% 0.02% 0.01% 0.02%
H2 0.01% 0.00% 0.01% 0.06%
SG (-) 0.617 0.626 0.727 0.689

Model Descriptions
PVT. Field gas sample compositions from two existing CDB wells, located in separate fields, A and B, were used to
model gas PVT behavior. PVTsim14 software was used to characterize the gas with SRK equation of state and to prepare the
PVT tables required for the wellbore dynamic simulations. The gas compositions from the two wells are shown in Table 2.
Two separate PVT tables, one for each field, were developed, given the significantly different H2S concentrations.
Phase Equilibium. Phase equilibrium modeling were derived from three sources. Alberta Sulphur Research Ltd.
(ASRL) generated elemental sulfur solubility phase diagrams using data from five CDB gas sour gas fields.11 Phase diagrams
were generated to estimate the solubility gradient for the sour gas under static conditions, as shown in
Fig. 4, for Field A. The near vertical dashed curve, on the right edge of the figure, is the melting point curve for
elemental sulfur at equilibrium with the respective H2S mixture.
Chevron Energy Technology Company (ETC) also developed a thermodynamic model for CDB sour gas.1 A PR-EoS was
developed to predict solubility of elemental sulfur in hydrocarbon gases. The property calculation module of HYSYS 2006
was the main platform for calculations. The critical properties reported by Reid et al.14 were used for calculations, as shown
in Table 3. The binary interaction parameters between all the components except that of S8 were taken from a Chevron
internal phase equilibrium model. The manual regression (trail and error) of interaction parameters between S8 and other
components indicated that the best fit was achieved by setting all S8 component BIPs to zero, except for S8-methane which
was set to 0.025. Table 4 shows the BIPs for the tuned PR-EoS. The sulfur solubility data used for the two CDB fields
are given in Table 5 and Table 6, resp.
The third method for S8 solubility was based on laboratory synthetic gas having the composition of CDB measured
surface gas from Field A. Table 7 shows the sulfur solubility data based on laboratory synthetic gas for CDB Field A gas.
Fig. 5 compares the sulfur solubility based on results from ASRL and ETC models, and the synthetic gas, for Field A.
Note the relatively close agreement between the ETC and synthetic gas results. Accordingly, the ETC sulfur phase
equilibrium model was chosen for the numerical modeling portion this study.
SPE 147244 5

Table 2: Gas compositions Table 3: Critical properties of gas Table 4: Binary interaction parameters for PR-EoS

Composition Mole (%) Tc (°F) Pc (psi) Omega MW C1 N2 CO2 H2S S8 C2 C3


Well A, Well B,
Field A Field B
C1 -116.4 673.1 0.011 16.04 C1 0.036 0.100 0.085 0.025 0.002 0.007
C1 82.08 74.61 N2 -232.5 492.3 0.04 28.01 N2 0.036 -0.020 0.165 0.000 0.050 0.085
C2 0.06 0.03 CO2 0.100 -0.020 0.097 0.000 0.130 0.135
CO2 87.7 1068.9 0.225 44.01
C3 0 0.02
H2S 9.13 15.97 H2S 212.8 1296.2 0.1 34.08 H2S 0.085 0.165 0.097 0.000 0.075 0.075
CO2 8.08 8.83 S8 1905.5 3002.3 0.4439 256.00 S8 0.025 0.000 0.000 0.000 0.000 0.000
N2 0.58 0.52
C2 90.1 708.3 0.0986 30.07 C2 0.002 0.050 0.130 0.075 0.000 0.001
He 0.02 0.01
H2 0.07 0.01 C3 206.1 617.4 0.1524 44.10 C3 0.007 0.085 0.135 0.075 0.000 0.001

Table 5: Chevron ETC S8 solubility, Field A Table 6: Chevron ETC S8 solubility,Field B Table 7: S8 solubility, synthetic gas

T P S 8 S ol u b i l i ty

3
℃ MP a g/ m

30 30 0. 0 8 9

30 40 0. 1 2

50 30 0. 1 1 8

50 40 0. 1 5 1

70 35 0. 1 5 7

70 40 0. 1 9 6

90 40 0. 2 7 6

90 45 0. 3 5 6

120 50 1. 1 8 9

OLGA Wax Module for Elemental Sulfur Deposition. There are presently, to the authors’ knowledge, no models
available in the industry to simulate and predict elemental sulfur precipitation and deposition in the tubing. We therefore
adapted the OLGA Wax Deposition Model6 (OWD) as a proxy. It is assumed that the sulfur deposition has same controlling
mechanisms as the wax deposition.
The OWD model is also a dynamic heat model, predicting transient temperature and pressure in the bulk gas flow stream
and the tubing wall, based on heat capacity and thermal conductivity properties of the gas, well components, and rock, and on
the geometry of the tubing.
The OWD model predicts precipitation in both the bulk gas flow stream, and deposition on the tubing wall, by comparing
bulk stream and tubing wall conditions (pressure and temperature), separately, relative to saturation conditions. The
precipitation rate in the bulk stream is assumed governed by:

C C C T
φ M M                      (1) 
T

The precipitated particles travel with the gas and do not deposit on the tubing wall, if wall conditions are not saturated.
6 SPE 147244

Sulfur Solubility  Comparison:  ASRI/ETC/Lab  Test Data 


0.8

90°C (ASRL)
0.7

70°C (ASRL)
Sulfur Solubility,  g/m3 0.6

90°C (ETC)
0.5

0.4
90°C (Lab)

0.3

0.2 70°C (ETC)
70°C (Lab)

30°C (Lab)
0.1

30°C (ASRL)
0
0 5 10 15 20 25 30 35 40 45 50

Pressure, MPa

Fig. 5 Sulfur solubility for Field A gas: ASRL / ETC / synthetic gas

Sulfur deposits on the wall when wall conditions become saturated. The deposition rate is controlled by molecular
diffusion through the laminar sub-layer of the fluid mass concentration profile. The diffusion rate is proportional to the
difference in sulfur solubility at bulk stream conditions and wall conditions. The rate of deposition per meter of pipe is
estimated by

D C C
N                               (2) 

Reference 6 provides further discussion on the different correlations that have been used for estimating the laminar sub-
layer thickness of mass concentration.
We used the framework of ODW for S8 deposition. The wax table is substituted with a sulfur solubility table with P/T.
The solubility curves in the sulfur table were based on the solubility curves from the Chevron ETC’s study. The code was
developed also to generate other parameters such as sulfur densities, molecular weight, enthalpies, thermal capacities and
thermal conductivities at various pressures and temperatures. The wax mass conservation equations are used to solve for
masses of sulfur dissolved in the gas, mass sulfur precipitated in the gas, and mass of sulfur deposited on the wall.

Basic Data
Well A was chosen to represent Field A, regarding gas composition and well geometry, and similarly, Well B to represent
Field B. The well trajectories are given in Fig. 6. Note Well B is significantly deeper and longer than Well A.
Well inflow performance is estimated from pressure transient test bottomhole data, corroborated with separate well
deliverability surface test data. Fig. 7 presents the resulting inflow performance model and tubing curves. Note Well A
deliverability is significantly higher than that of Well B.
SPE 147244 7

1000

2000

3000
TVD, m

Well A
4000

5000

Well B
6000

7000
0 500 1000 1500 2000 2500

Horizontal Displacement, m

Fig. 6 Well trajectories (Well A and WellL B)

IPR/VLP Plots: Well A and Well B


10000

IPR: Well A
VLP: Well A Tubing
8000 IPR, Well B
VLP- Well B Tubing

6000
BHP, psia

4000

2000

0
0 300 600 900 1200 1500
Gas Rate, MMscf/d

Fig. 7 Inflow and outflow curves: WELL A and Well B


8 SPE 147244

The casing and tubing dimensions and configurations are summarized in

Fig. 8 WELL A Configuration (proposed salvaging plan)

. Fig. 8 describes the salvaged well configuration for Well A in Field A. The existing wellbore in Well A has the risk
from predominantly non-CRA casing used in sour gas service and poor cement bond. The salvage plan is to run 7” x 4-1/2”
new string and cement to surface in the annulus between the new string and existing string, with all CRA tubulars, except in
the 7 in. above the upper packer.
Geothermal relations, in oC, are the following:

T 0.021 TVDKB 16.7 (Well A) (3)

T 0.0198 TVDKB 23.3 (Well B) (4)


SPE 147244 9

Table 8: Well configuration

D e pt h, m
S t ring O D , in
We ll A We ll B

Structural P ipe 20 0-31.7 0 – 60

Surface Casing 13.375 0-470.3 0 – 1000

Technical Casing 9.625 0-2857.5 0 – 4550

Casing 7 1841.9-3598.5 0 – 6285

P DL 5.5 3598.5-4142.1 0 – 6821

Tubing 4.5 0-3599 0 - 6285

Fig. 8 WELL A Configuration (proposed salvaging plan)


10 SPE 147244

Results and Discussion


Steady State Estimations
A general understanding of S8 precipitation behavior, and precipitation rates relative to tubing gas flow velocity, provide
insight into the flow assurance risk presented by S8 deposition. Accordingly, preliminary deposition estimations were
generated using more simple flow models that predict steady state pressure and temperature variation in the tubing at
expected operating well rates.
The steady state estimations allow for a basic understanding of S8 precipitation behavior for Field A. Recalling Fig. 4,
note the pressure and temperature paths for Well A, at well rates of 50 MMscf/d and 10 MMscf/d. These paths were
estimated using the Petroleum Experts IPM Prosper Model. Three points in the paths are indicated – reservoir, wellhead, and
field pipeline (“flowline”).
Note that 80% of the total precipitation occurs in the well - the saturation concentration decreases from 0.5 g/m3 at
reservoir conditions, to 0.1 g/m3 at wellhead conditions. Therefore, the demand on well velocity to suspend sulfur precipitate
is significant, although not as great as the demand on surface facilities to keep precipitate moving. This demand consists of
the precipitate exiting the well, and an additional 20% precipitating through the well choke and gathering lines.
The importance of the uncertainty in reservoir in-situ S8 concentration is also seen in Fig. 4. Reservoir S8 concentrations
are not shown here, only saturation concentrations estimated from the ETC phase equilibrium model. 80% of total
precipitation occurs in the well, but only if the in-situ S8 concentration is initially 0.5 g/m3, coincidentally equal to the
estimated saturation concentration. If the in-situ concentration is instead, for example, less than 0.1 g/m3, there will be no
precipitation at the well. Additionally, a significantly lower burden is placed on the surface lines, as only 0.1 g/m3 will be
required to be carried downstream of the well, compared to almost 0.5 g/m3 in the saturated case. If the reservoir in-situ
concentration is less than 0.5 g/m3, the reservoir is said to be undersaturated in S8.
The steady state flow results also show precipitation rates along the flow path in the tubing, compared to gas velocity
profiles in the tubing.
These profile comparisons are estimated for various saturation levels, and using both ETC and ARSL S8 phase models for
comparison. Here, saturation level is defined as the ratio of saturation pressure to reservoir pressure, in percent. This
parameter captures the state of undersaturation in terms of pressure, rather than concentration, as discussed above. For
example, if reservoir pressure and saturation pressure are equal, the saturation level is 100% (and the in-situ concentration is
0.5 g/m3). If alternatively, saturation pressure is lower than reservoir pressure, that is, the reservoir is undersaturated in S8,
the saturation level is less than 100% (and the in-situ concentration is less than 0.5 g/m3) .
Fig. 9 and 10, and Table 9 and 10, show sulfur precipitation rate, in lbs/day, as it accumulates from reservoir depth to
wellhead, in Well A tubing. The model rates are 10 and 50 MMscf/d, and both ETC and ASRL phase model results are
shown for the 50 MMscf/d case, in Fig. 10, for comparison. Note the ASRL phase model precipitation rate predictions
exceed those of the ETC model, by 50%. The results for various S8 saturation levels are indicated.
Note the comparison in precipitation rates between the two gas well flow rate cases indicate nearly direct proportionality
with rate. Thus the differences in flowing pressure and temperature between the two gas flow rates are not enough to cause
significant differences in saturation concentrations. This result can also be seen in Fig. 4, where the predicted saturation
concentrations are similar for each gas rate at the wellhead and pipeline.

Sulfur Precipitation Rate,  lbs/day
0 50 100 150 200 250 300 350
0

500

1000

1500 S8 100% Saturated
S8 75% Saturated
2000 S8 50% Saturated
MD, m

2500

Tubing Velocity:   4 ‐ 5 ft/sec
3000
Flowline velocity:  8 ft/sec

3500

4000

Fig. 9 Well A cumulative S8 precipitation vs. depth at 10 MMscf/d Fig. 10 Well A cumulative S8 precipitation vs. depth at 50 MMscf/d

Table 9 indicates that for Well A at 50 MMscf/d, and S8 saturation level of 100%, a cumulative 1200 lb/day sulfur will
precipitate along the entire length of the tubing. The estimated superficial steady state gas velocity at the wellhead in this
case is 28 ft/sec, which has to be sufficient to carry this estimated mass to avoid precipitation migration to the well bottom.
SPE 147244 11

Further, the all gas velocities in the table, corresponding to various depths, must be sufficient to carry the associated
cumulative precipitate rate at that depth.
Note that an S8 saturation level of 50% results in no precipitation in the wellbore.
If the well produces at 10 MMscf/d, 100% S8 saturation level generates 270 lb/day sulfur in the well, as shown Table 10.
The calculated superficial gas velocity is 4.5 ft/sec at the wellhead, which has to be sufficient to suspend this amount of mass
to avoid botomhole accumulation.

Sulfur Precipitation Rate,  lbs/day
0 50 100 150 200 250 300 350
0

500

1000

1500 S8 100% Saturated
S8 75% Saturated
2000 S8 50% Saturated
MD, m

2500

Tubing Velocity:   4 ‐ 5 ft/sec
3000
Flowline velocity:  8 ft/sec

3500

4000

Fig. 9 Well A cumulative S8 precipitation vs. depth at 10 MMscf/d Fig. 10 Well A cumulative S8 precipitation vs. depth at 50 MMscf/d

Table 9: S8 precipitation prediction: Well A, 50 MMscf/d, ASRL phase model

4‐1/2 tbg, 50 MMSCFD

100% Saturated 75% Saturated 50% Saturated


Gas 
MD  P T
Velocity
Sat. Precipitation  Sat. Precipitation  Sat. Precipitation 

m psia F g/m3 kg / d lb / d g/m3 kg / d lb / d g/m3 kg / d lb / d ft/s
Choke 1315 88 0 687 1511 0 304 670 0 64 140 82
0 4027 160 0.11 531 1168 0.11 149 327 0.045 0 0 27.9
1200 4600 174 0.185 425 934 0.185 42 93 0.045 0 0 26.2
2500 5260 185 0.34 205 452 0.215 0 0 0.045 0 0 24.4
3595 5761 189 0.475 14 31 0.215 0 0 0.045 0 0 23.1
3695 5838 189 0.485 0 0 0.215 0 0 0.045 0 0 16.5

Table 10: S8 precipitation prediction: Well A, 10 MMscf/d, ASRL phase model

4‐1/2 tbg, 10 MMSCFD
100% Saturated 75% Saturated 50% Saturated
Gas 
MD  P T
Velocity
Sat. Precipitation  Sat. Precipitation  Sat. Precipitation 

m psia F g/m3 kg / d lb / d g/m3 kg / d lb / d g/m3 kg / d lb / d ft/s
Choke 1315 24 0 137 302 0 61 134 0 13 28 14.2
0 4506 104 0.055 122 268 0.055 45 100 0.045 0 0 4.5
1200 4964 142 0.155 93 206 0.155 17 37 0.045 0 0 4.6
2500 5462 176 0.35 38 84 0.215 0 0 0.045 0 0 4.7
3595 5826 189 0.482 1 2 0.215 0 0 0.045 0 0 4.6
3695 5838 189 0.485 0 0 0.215 0 0 0.045 0 0 3.3

Deposition of S8 in the Tubing


S8 precipitation predictions using steady state flow models can not address prediction of deposition, whether through
accumulation at the well bottom, or deposition on the tubing wall. The OWD model is employed for these predictions. The
phase equilibrium model chosen to be used in conjunction with the OWD model is the ETC model, given its consistency with
the synthetic gas solubility results.
12 SPE 147244

Base Case: Well A, 50 MMscf/d, 4.5 in. Tubing. A gas rate of 50 MMscf/d and a tubing OD of 4.5” (ID 3.958”) were used
as the base case sulfur deposition simulation for Well A. A 60-day flow at 50 MMscf/d was modeled, as shown in Fig. 11.
The increase in the deposition thickness at 15-day intervals is presented. Note that for the OWD model, the tubing length
initial point is the well bottom.
The maximum sulfur deposition occurs just upstream of the wellhead choke location, which is about 0.54 mm, at 60 days.
Beyond 60 days, 3.9 years is required to form a layer of 0.5-in. thickness. The mass rate of the sulfur production in this case
is 500 kg/day at the wellhead.
g
0.6
E (WH choke)
Sulfur Deposition Thickness,  mm

0.5
C (4”x7” cross‐over)
0.4
B (7” csg shoe) D (surface casing)
15 days
0.3
30 days
0.2 45 days
A (above PDL)
60 days
0.1
BH
0
0 500 1000 1500 2000 2500 3000 3500 4000 4500

Well MD from Bottom to Wellhead,  m

Fig. 11 S8 precipitation during production: Well A, 50 MMscf/d, 4.5-in. tubing

Keep in mind that the deposition rate is controlled by the molecular diffusion through the laminar sub-layer of the fluid
velocity profile. The diffusion rate is proportional to the difference in sulfur solubility at the bulk temperature and at the
tubing wall surface temperature. Considerating the specific tubing/casing/cement insulation and configuration as shown in
Fig. 8, the following gives some interpretation of the simulation results with a varying S8 deposition profile along the
wellbore.
• Note that there is 544 m wellbore lateral section where no temperature difference exists between the inflowing
gas stream and the PDL wall in the reservoir contact. Thus the S8 deposition is zero.
• The S8 depostion starts from tubing at 800 m (Point A) above the bottomhole, and the S8 deposition trend
increases from 800 m to 1250 m due to increasing temperature difference in the upward direction between the
bulk gas temperature and wall temperature. The increasing differene in temperature comes from the fact that the
ambient temperature or geothermal temperature decreases faster upward than the cooling-down speed of the hot
gas stream in the tubing.
• The increasing trend of S8 deposition terminates at 1250 m (Point B) where another thermal layer by the 7-in
technical casing results in less heat conduction and thus the less temperature difference between the gas stream
and the tubing wall due to improved thermal insulation.
• A new temperature difference trend builds up above the 7-in casing shoe position until 2250 m (Point C) where
the thick cement insulation ends at the 4"x7" cross-over.
• From 2250 m upward (point C), the overall heat transfer becomes faster due to less insulation layer around the
tubing. Thus the higher temperature difference exists between the gas stream and the tubing wall due to poorer
thermal insulation.
• In the last 200 m section, the maximum sulfur deposition occurs just upstream of the wellhead choke location in
the model. The quickly increased gas velocity is more dominant than the temperature difference change along
the upward direction. The significantly reduced boundary layer increases the S8 concentration gradient in the
diffusivity equation (Eq. 2).

A 500-day shut-in, following the 60-day flow, was also modeled. This duration of shut-in is sufficient for the well to cool
to the original geothermal state, as shown in Fig. 12. The thickness of sulfur deposited at the wall after the shut-in is the same
as reported at the end of the flow period, indicating no additional sulfur deposition on the tubing wall.
SPE 147244 13

The majority of precipitate accumulates in the well bottom during the shut-in. At the end of the flow period, the mass of
suspended sulfur in the tubing is 1.4 kg, as shown in Fig. 13. During the shut-in, as the gas cools, an additional 0.4 kg
precipitates. The total, 1.8 kg, falls to the well bottom.

TVD from the Bottomhole, m

Well MD from Bottom to Wellhead

Fig. 12. S8 deposition after 500-day shut-in, Well A

Fig. 13 Mass of sulfur precipitation during 500 days of shut-in, WELL A well base case

Mechanism of S8 Deposition in Tubing. Based on the S8 deposition modeling results, it indicates two types of sulfur
deposition mechanisms occur in the tubing during production and shut-in. The first mechanism is deposition on the tubing
wall, of precipitate from the gas flow boundary layer adjacent to the wall. This occurs primarily during production, at a slow
deposition rate. The deposition rate of this mechanism is insignificant during shut-in. This is due to the lack of new gas
entering the well after shut-in. The S8 mass available for precipitation is therefore restricted to that of a single tubing volume.
Additionally, deposition rate decreases because the gas cools during the shut-in, resulting in a decrease in S8 concentration
gradient near the tubing wall.
The second mechanism is deposition of S8 precipitate from the bulk flow stream at the bottomhole. This occurs primarily
during shut-in. The rate of deposition for this mechanism during flow periods is insignifcant, since, at nominal gas flow
rates, the flow velocity exceeds the precipitate particle falling velocity (critical velocity).
These two mechanisms are illustrated in Figures 14 and 15.
Assuming a sulfur particle diameter of 10 microns, the slip velocity in the gas is estimated at 0.0019 m/s using Stokes’
Law. Based on this slip velocity calculation, the accumulation at the horizontal well bottom, more precisely at the angle build
section for the lateral, of all the sulfur particles takes almost 21 days during the shut-in.
This time estimate may be high as it assumes no particle coalescence during the fall. If a sulfur particle is assumed to
have a diameter of 100 microns, fior example, the slip velocity is estimated to be 0.052 m/s, 27 times greater than the 10
14 SPE 147244

micron particle slip velocity. At the higher slip velocity, the accumulation of the entire sulfur precipitate mass at the angle
build section occurs in 19 hours.
Presently, no published studies exist, to the authors’ knowledge, as to S8 particle size distribution under operating
conditions, and so the uncertainty in precipitate accumulation time at the angle build section, is significant.
During production, the uncertainty in particle size has little relevance, as estimated gas production superficial velocities,
at nominal gas production rates, are much higher than estimated critical velocities for large particles (see gas velocities for
10 MMscf/d and 50 MMscf/d in Tables 9 and 10).
The geometry of the accumulation at the angle build section is highly uncertain, as the OWD model does not predict such.
If the S8 uniformly fills the tubing, a typical 1.8 kg mass of sulfur (see following section) would occupy a length of 11 cm
inside the 4.5 in. tubing (3.958 in. ID). In a high-angle well, the deposit may occupy a longer interval along the low side of
the hole.
At the resumption of production, all or part of the accumulated sulfur might be removed by the flow. There may be a risk
that the deposit may bond to the tubing. Under this bonding hupothesisis, several shut-ins and sulfur accumulations at the
same location in the tubing may result in a flow restriction.

Fig. 14. Primary S8 deposition mechanism during production Fig. 15. Primary S8 deposition mechanism during shut-in

Well A, 10 MMscf/d, 4.5 in. Tubing. Sulfur deposition thickness on the tubing wall at 60 days of flow at a lower gas rate of
10 MMscf/d, is shown in Fig. 16. The maximum sulfur deposition is about 0.31 mm. Unlike at the higher rate of 50
MMscf/d, for which the maximum accumulation thickness occurred at the wellhead, the maximum deposit thickness for this
lower rate case is predicted to occur deeper, at nearly the lower part of the tubing length where the gas stream drag force
might be less due to lower gas velocity thus S8 particle is easier to deposite on the lower tubing wall than at the upper tubing
section. Assuming a constant deposition rate at that location, it takes about 6.8 years to form 0.5-in. layer. The total mass rate
of the sulfur precipitation in the tubing is 90 kg/day at this gas rate.
Note that generally the predicted deposit thickness for this lower rate is less than that of the 50 MMscf/d case. Fig. 17
presents a comparison between the two flow rate cases, showing tubing wall deposit thickness, and tubing temperature.
SPE 147244 15

0.35

0.3

Sulfur Deposition Thickness,  mm
0.25

0.2
15 days

0.15 30 days
45 days
0.1 60 days

0.05

0
0 500 1000 1500 2000 2500 3000 3500 4000 4500

Well MD from Bottom to Wellhead,  m

Fig. 16 S8 precipitation during production: 10 MMscf/d, 4.5-in. tubing, WELL A

Well MD from Bottom to Wellhead

Fig. 17 Sulfur deposition comparison between 50 MMscf/d and 10 MMscfd gas rates, Well A

The reason for the less S8 thickness is the lower sulfur in solution in the cooler bulk flow stream at this lower rate. Note
in Fig. 17, the lower predicted temperature in the tubing for the lower rate. The lower concemtration in the bulk stream
results in a smaller concentration gradient between the bulk stream and the boundary layer, and thus a smaller S8 diffusion
rate. Accordingly, a smaller deposit rate on the tubing wall results.
For both rates of 50 MMscf/d and 10 MMscf/d, there are thre S8 deposition humps located in the same wellbore locations,
which is believed to be basically related to the overall wall heat transfer coefficient, as explained in Fig. 11. Fig. 18 compares
the overall heat transfer coefficient (Q2 in OLGA variable) and S8 deposition thickness (DXWX in OLGA variable) plots for
the 50 MMscf/dand 10 MMscf/d. It can be found that the three deposition locations are where the overall heat transfer
coefficient make a step change.
16 SPE 147244

Well MD from Bottom to Wellhead

Fig. 18 Sulfur deposition comparison between 50 MMscf/d and 10 MMscfd gas rates, Well A
 
A 500-day shut-in was modeled which followed the 60-day flow period. An insignificant accumulation on the tubing
wall was predicted during the shut-in, as was predicted for the shut-in following the higher rate flow.
At the end of the flow period, the precipitated mass in suspension was 1.5 kg, slightly higher than the 50 MMscf/d rate
due to the lower bulk stream temperature and therefore lower S8 saturation. The shut-in adds an additional 0.3 kg of sulfur
precipitate. The final mass of suspended sulfur is 1.8 kg. This is equivalent to the final precipitated mass in the 50 MMscf/d
case, as at the start of the shut-in, the density of gas in the tubing is similar – the tubing pressure and temperature for the two
cases at start of shut-in are similar.

Well A, 50 MMscf/d, 3.5 in. Tubing. An alternative completion was proposed for Well A in which 3.5 in. (OD) tubing
replaced 4.5 in. (OD) tubing in the upper 1800m portion of the well. Well A is a well to be salvaged by using 4.5” monobore
at the lower section, with a crossover to connect to the upper tubing. In the salvage well configuration, the simulated 3.5”
tubing is thus only for the 1800 m upper tubing (see Fig. 8). A comparison between sulfur accumulation thickness, and tubing
wall temperature, for the two tubing cases is presented in Fig. 19.
The predicted bulk stream temperature in 3.5 in. tubing is lower than in 4.5 in. This is due to larger pressure drop with
smaller tubing and thus larger J-T cooling. The sulfur deposition rate in 3.5 in. tubing is therefore less, since, as previously
explained, the concentration gradient between the bulk stream and bounadary is smaller. Lower gas temperature in the upper
tubing results in less S8 deposition, thus the S8 deposition rate can not recovered even at the wellhead to the same level at the
cross-over location.
The maximum sulfur deposition of 0.4 mm after the 60-day flow occurs around 1800 m MD from the wellhead.Assuming
a constant deposition rate, a layer of 0.5-in. requires 4 years to accumulate.

Well MD from Bottom to Wellhead

Fig. 19 Sulfur deposition comparison with 4.5 and 3.5 in OD tubing at 50 MMscf/d, Well A
SPE 147244 17

Shut-in simulation following the flow period again results in insignificant presicted additional tubing wall accumulation.
At the end of the 60-day flow, the sulfur mass precipitated is 1 kg. An 0.28 kg of additional sulfur precipitates during the
shut-in. The final mass of suspended sulfur is 1.28 kg, which is consistent with the smaller tubing volume of the alternative
completion.

Well B, 30 MMscf/d, 4.5 in. Tubing. Well B, located in Field B, has a lower reservoir deliverability than Well A, and
accordingly, a lower predicted operating rate. Recall however, that the reservoir is deeper and hotter in Well B. Potentially,
the reservoir in-situ S8 concentration is higher than at Well A (compare Tables 5 and 6), and so the sulfur deposition in the
well may be higher.
The predicted deposition profile at the end of a 60-day flow, of this deeper, hotter well is significantly different than at
Well A, as seen in Fig. 20. The bottom portion of the well, some 2300m, is predicted to have little or no deposition. This is
due to a relatively small temperature drop over this interval (temperature is still high there) due to its high angle (see Fig. 6).
The maximum deposition thickness is 2.43 mm, significantly greater than at Well B, and occurs at 835 m from the
wellhead. This is due to that Well B has higher S8 saturation capacity and lower operating rate and colder wellhead
temperature than Well A.
Assuming a constant deposition rate, 0.9 years is required to accumulate a 0.5-in. layer, which is a rate that may be
significantly more problematic than at Well A.
The predicited total mass rate of sulfur precipitation which does not deposit on the tubing wall, is 950 kg/day, nearly
twice that predicted for Well A. Thus, both tubing deposition, and surface facility deposition, may be significantly higher at
Field B than at Field A.

3
Sulfur Deposition Thickness,  mm

2.5

15 days
1.5
30 days
1 45 days
60 days
0.5

0
0 1000 2000 3000 4000 5000 6000 7000

Well MD from Bottom to Wellhead,  m

Fig. 20 S8 precipitation during production: 30 MMscf/d, 4.5-in. tubing, Well B

The shut-in precipitation prediction for Well B is also significantly different than for Well A. The mass of precipitated
sulfur is 5 kg at the end of the 60-day flow, exceeding 3 times that for Well A. This is due to the increase in both the S8
saturation, and tubing volume in Well B. During the 500-day shut-in, 3.8 kg of additional sulfur precipitates, exceeding 12
times that predicted for Well A. The total sulfur precipitate mass is 8.8 kg. Assuming a sulfur particle diameter of 10
microns, the corresponding slip velocity allows the sulfur to accumulate at the angle build section in 36 days. Assuming a
sulfur particle diameter of 100 microns, the accumulation time is reduced to 33 hours. 8.8 kg of sulfur would form a 54 cm
plug in the tubing, assuming a uniform fill, five times longer than predicted for Well A .
Field B is assumed to be at higher risk for flow assurance in the presence of S8 production, based on these results. Wells
in Field B may therefore require more frequent, and larger, solvent treatments to mitigate S8 deposition in the tubing.

Comparison and Summary: Table 11summarizes the OWD model results discussed above. Following are the principal
observations from the OWD modeling.
• Tubing wall deposition is likely to occur during flow periods only. S8 deposition on the tubing wall accumulates
gradually, and is therefore likely amenable to mitigation by regular solvent treatment.
• Botomhole accumulation is likely to occur during shut-in periods only.
• Tubing wall deposition is correlated positively to gas well rate. Higher operating rate wells may therefore require
more frequent sulfur solvent treatment.
• Higher gas well rate tubing wall depositions occur nearer the wellhead, while lower rate depositions occur deeper.
18 SPE 147244

This result may influence the solvent treatment procedure.


• Bottomhole accumulation rate during shut-in periods is strongly dependent on S8 particle size distribution, which
presently is highly uncertain.
• The accumulation geometry is highly uncertain and is not predicted in this stud. In both wells A and B, the
accumulation section of the tubing is likely to be the angle build section for the lateral.
• Well A modeling indicates that it is unlikely that significant S8 accumulates on the tubing wall, or in the well
bottom, to cause restriction to flow, in Field A.
• Well B modeling indicates S8 accumulation on the tubing wall, or at the well bottom, may cause restriction to flow,
and could occur within one year of continuous production, in Field B.
• Field B wells and surface facilities may have to carry twice the mass of sulfur as in Field A.
• For Field B, although lower rate has slower S8 depostion rate during production, the S8 accumulation is higher
during shut-in. Thus lower rate of production is not recommended.
• The tubing sizes proposed at CDB do not result in significantly different S8 deposition predictions.

Table 11: Summary of Sulfur deposition prediction based on dynamic simulation

Thickest S8  S8 mass  Time for S8  Time for S8 


Time to  Maximum  S8 plug 
S8 thickness  depostion  rate  Suspended  particles to  particles 
form 0.5"  S8 falling to  length 
Field /  Well  Qg,  Tubing  on tubing  location,  transported  S8 in  fall to  falling to 
Case S8 layer  bottom  invertical 
Well TVD, m MMscfd OD, in. wall in 2  MD from  to surface  steady  bottom ( 10  bottom (100 
on tubing  hole in  section, 
months, mm well head,  in steady  state, kg microns),  microns), 
wall, year shut‐in, kg mm
m state, kg/d day day
Base 
Case 50 4.5 0.54 0 m 3.9 466 1.4 1.8 11.0
Field B /  Case 1 10 4.5 0.31 2850 m 6.8 90 1.5 1.82 11.1
3506 21 0.8
Well B
Case 2 50 3.5 0.4 1800 m 5.3 466 1 1.28 12.9
Case 3 10 3.5 0.32 2850 m  6.6 90 1 1.29 13.0
Case 4 30 4.5 2.43 835 m 0.9 950 5 8.8 54.0
Field A /  Case 5 5 4.5 0.64 4300 m 3.3 188 9 10.8 66.3
6000 36 1.4
Well A Case 6 30 3.5 1.7 835 m 1.2 950 3 5.2 52.7
Case 7 5 3.5 0.67 4300 m 3.2 188 5.4 6.5 65.9

Conclusions
A flow assurance study regarding elemental sulfur deposition in production tubing has been performed for the CDB sour gas
project. Elemental sulfur dissolved in the sour gas stream may precipitate and deposit on the tubing and at the well bottom,
and may cause flow restrictions. Flow restrictions may also occur in surface facilities, as observed at an offset field having
similar gas properties, although surface deposition is not a subject of this study.
Lack of bottomhole samples at CDB leave the reservoir gas in-situ S8 concentration as highly uncertain. The dependence
of sulfur precipitation rates on in-situ concentration is presented. However, a conservative approach is taken in which, for the
tubing wall and well bottom deposition rate and mass modeling, the reservoir gas is assumed to be fully saturated with
S8.Bottomhole sampling is planned at CDB to reduce the in-situ S8 concentration uncertainty.
Significant sulfur precipitation may occur in the tubing, relative to the entire flow system, but is likely fully carried to the
surface during flow periods, even under high S8 saturation cases.
Based on the simulation, surfur deposition mechanisms in the wellbore are presented. Flow assurance within the well is
risked by growth of S8 on the tubing wall, during production. During shut-in, flow assurance is risked by accumulation at the
well bottom, but the risk is strongly dependent on S8 particle size distribution, with larger particles accumulating at a greater
rate, and thereby possibly forming restrictions during the limited time of the shut-in. The risk is also strongly dependent on
the geometry of the accumulation, with accumulation at the angle build section of horizontal laterals, possibly laying on the
low side of the hole and thereby may not significantly restricting gas well flow. To be safe, it is still suggested that bull
heading S8 solvent is always required during shut-in.
Specifically regarding the CDB wells modeled, principal conclusions are as stated in the previous section.

Nomenclature
C = coefficient for gas back-pressure equation, MMscf/d/psi2
C sulfur solubility at gas flow bulk temperature, moles/m3
C sulfur solubility at wall surface temperature, moles/m3
C =elemental sulfur concentration, kg/kg
D diffusion coefficient of sulfur, m2/s
M = gas mass, kg
N rate of deposition on wall, moles/ m s
SPE 147244 19

P = pressure, Pa
r pipe radius, m
t = time, s.
T = temperature, ºC
SIWHP = shut-in wellhead pressure, psia
FWHP = flowing wellhead pressure, psia

Greek symbols
φ = precipitation rate, kg/s
δ laminar sub‐layer thickness of mass concentration, m.

Acknowledgments
The authors would like to thank Chevron Asia South E&P and Chuandongbei Gas Project for permission to publish this
study. Special thanks to following people: Hussein Alboudwarej (Chevron ETC), Dallas Thill (Chevron ASBU), Robert
Marriott and Paul Davis (ASRL), Honggang Chang and Meng Tang (Petro-China RINGT), Mike Ye and Chaohong Xiao
(Chevron Chuandongbei gas project) for their discussion, suggestion and contribution to this study. We appreciate Jerome
Glass of Chuandongbei lead development geologist for his review of this paper.

References
1. Alboudwarej, H.: “Sulfur Solubility Study for Gas Samples from CDB Field (China)”, Chevron Energy Technology Company, San
Ramon, CA, U.S.A., May 23rd, 2008, Technical Report.
2. Bendiksen, K.H., Malnes, D., Moe, R. and Nuland, S.: “The Dynamic Two-Fluid Model OLGA: Theory and Application,” SPE
Production Engineering, May 1991, pp. 171-180
3. Brunner, E. and Woll, W.: “Solubility of Sulfur in Hydrogen Sulfide and Sour Gases,” Society of Petroleum Engineers Journal,
1980(5): p. 377-384.
4. Brunner, E., et. al.: “Sulfur solubility in sour gas,” Journal of Petroleum Technology, 1988. 40(12): p. 1587-1592.
5. Heidemann, R.A., et al., “A Chemical Equilibrium Equation of State Model for Elemental Sulfur and Sulfur-Containing Fluids,”
Industrial and Engineering Chemistry Research, 2001. 40(9): p. 2160-2167.
6. Hovden, Lars et. al.: “Pipeline Wax Deposition Models and Model for Removal of Wax by Pigging: Comparison between Model
Predictions and Operational Experience,” Conference Publications Proceedings, the 4th North American Conference on Multiphase
Technology Banff, Canada, 3-4 June 2004
7. Karan, K., Heidemann, R.A. and Behie, L.A.: “Sulfur Solubility in Sour Gas: Predictions with an Equation of State Model,” Industrial
& Engineering Chemistry Research, 1998. 37(5): p. 1679-1684.
8. Keskin, C: “Dynamic Simulation Study on Sulfur Deposition in Wellbore for Chevron China CDB Sour Gas Project,” SPT Group,
Houston, TX, U.S.A., December 27, 2010
9. Marriott, R. A.: “Progress towards Predicting Elemental Sulfur Solubility In High Pressure Sour Gas” , ASRL Sour Gas Prod. Symp.,
Calgary, Alberta, 2008.
10. Marriott, R. A., Davis, P. M. and Clark, P. D. “Towards Estimating and Alleviating Sulfur Deposition during the Production Of Sour
Gas Wells” June ASRL Chalk Talks, 2005.
11. Marriott, R.A., Davis, P. M. and Clark, P. D. “Elemental Sulfur Phase Diagrams of Five Chuandongbei Sour Fluids”, ASRL #1164.01-
2008-2009, Calgary, Alberta , April 8th, 2010, Technical Report.
12. OLGA 5.3 User Manual, SPT Group, Aug., 2010
13. PVTSim User Manual (Ver. 19), Calsep A/S, 2010
14. Reid, R.C., Prausnitz, J.M., Poling, B.E.: “The Properties of Gases and Liquids,” McGraw-Hill: New York, 1987.
15. Wen, Y.C.: “Elemental Sulfur in Chuangdongbei Trassic Feixianguan formation,”Dec., 2008, , Technical Report.

You might also like