Chapter 4

Wellbore Flow Performance

Abstract

Wellbore performance analysis involves establishing a relationship among tubular size, wellhead and bottom-hole pressure, fluid properties, and fluid production rate. Understanding wellbore flow performance is vitally important to production engineers for designing oil well equipment and optimizing well production conditions. Oil can be produced through tubing, casing, or both in an oil well, depending on which flow path provides better performance. Producing oil through tubing is a better option in most cases to take the advantage of gas-lift effects. The mathematical models are also valid for casing flow and casing-tubing annular flow as long as hydraulic diameter is used. Among many models, the modified Hagedorn–Brown model has been found to give results with good accuracy. The industry practice is to conduct a flow gradient (FG) survey to measure the flowing pressures along the tubing string. The FG data are then employed to validate one of the models and tune the model if necessary before the model is used on a large scale.

Keywords

TPR; VLP; wellbore; flow; hydraulics

4.1 Introduction

Chapter 3, Reservoir Deliverability described reservoir deliverability. However, the achievable oil production rate from a well is determined by wellhead pressure and the flow performance of production string; that is, tubing, casing, or both. The flow performance of production string depends on geometries of the production string and properties of fluids being produced. The fluids in oil wells include oil, water, gas, and sand. Wellbore performance analysis involves establishing a relationship between tubular size, wellhead and bottom-hole pressure, fluid properties, and fluid production rate. Understanding wellbore flow performance is vitally important to production engineers for designing oil well equipment and optimizing well production conditions.

Oil can be produced through tubing, casing, or both in an oil well, depending on which flow path has better performance. Producing oil through ubing is a better option in most cases to take the advantage of gas-lift effect. The traditional term tubing performance relationship (TPR) is used in this book (other terms such as vertical lift performance have been used in the literature). However, the mathematical models are also valid for casing flow and casing-tubing annular flow as long as hydraulic diameter is used. This chapter focuses on determination of TPR and pressure traverse along the well string. Both single-phase and multiphase fluids are considered. Calculation examples are illustrated with hand calculations and computer spreadsheets that are provided with this book.

4.2 Single-Phase Liquid Flow

Single-phase liquid flow exists in an oil well only when the wellhead pressure is above the bubble-point pressure of the oil, which is usually not a reality. However, it is convenient to start from single-phase liquid for establishing the concept of fluid flow in oil wells where multiphase flow usually dominates.

Consider a fluid flowing from point 1 to point 2 in a tubing string of length L and height Δz (Fig. 4.1). The first law of thermodynamics yields the following equation for pressure drop:

ΔP=P1P2=ggcρΔz+ρ2gcΔu2+2fFρu2LgcD (4.1)

image (4.1)

where

ΔP=pressure drop, lbf/ft2

P1=pressure at point 1, lbf/ft2

P2=pressure at point 2, lbf/ft2

g=gravitational acceleration, 32.17 ft/s2

gc=unit conversion factor, 32.17 lbm-ft/lbf-s2

ρ=fluid density, lbm/ft3

Δz=elevation increase, ft

u=fluid velocity, ft/s

fF=Fanning friction factor

L=tubing length, ft

D=tubing inner diameter, ft

image
Figure 4.1 Flow along a tubing string.

The first, second, and third terms in the right-hand side of the equation represent pressure drops due to changes in elevation, kinetic energy, and friction, respectively.

The Fanning friction factor (fF) can be evaluated based on Reynolds number and relative roughness. Reynolds number is defined as the ratio of inertial force to viscous force. The Reynolds number is expressed in consistent units as

NRe=Duρμ (4.2)

image (4.2)

or in U.S. field units as

NRe=1.48qρdμ (4.3)

image (4.3)

where

NRe=Reynolds number

q=fluid flow rate, bbl/day

ρ=fluid density, lbm/ft3

d=tubing inner diameter, in.

μ=fluid viscosity, cp

For laminar flow where NRe<2000, the fF is inversely proportional to the Reynolds number, or

fF=16NRe (4.4)

image (4.4)

For turbulent flow where NRe>2100, the fF can be estimated using empirical correlations. Among numerous correlations developed by different investigators, Chen’s (1979) correlation has an explicit form and gives similar accuracy to the Colebrook–White equation (Gregory and Fogarasi, 1985) that was used for generating the friction factor chart used in the petroleum industry. Chen’s correlation takes the following form:

1fF=4×log{ε3.70655.0452NRelog[ε1.10982.8257+(7.149NRe)0.8981]} (4.5)

image (4.5)

where the relative roughness is defined as ε=δdimage, and δ is the absolute roughness of pipe wall.

The fF can also be obtained based on Darcy–Wiesbach friction factor shown in Fig. 4.2. The Darcy–Wiesbach friction factor is also referred to as the Moody friction factor (fM) in some literatures. The relation between the Moody and the fF is expressed as

fF=fM4. (4.6)

image (4.6)
image
Figure 4.2 Darcy—Wiesbach friction factor diagram. After Moody, 1944.

Example Problem 4.1 Suppose that 1000 bbl/day of 40 °API, 1.2 cp oil is being produced through 278image-in., 8.6-lbm/ft tubing in a well that is 15° from vertical. If the tubing wall relative roughness is 0.001, calculate the pressure drop over 1000 ft of tubing.

Solution Oil-specific gravity:

γ0=141.5°API+131.5=141.540+131.5=0.825

image

Oil density:

ρ=62.4γ0=(62.5)(0.825)=51.571bm/ft3

image

Elevation increase:

ΔZ=cos(α)L=cos(15)(1,000)=966ft

image

The 278image-in., 8.6-lbm/ft tubing has an inner diameter of 2.259 in. Therefore,

D=2.25912=0.188ft.

image

Fluid velocity can be calculated accordingly:

u=4qπD2=4(5.615)(1,000)π(0.188)2(86,400)=2.34ft/s.

image

Reynolds number:

NRe=1.48qpdμ=1.48(1000)(51.57)(2.259)(1.2)=28,115>2100,turbulentflow

image

Chen’s correlation gives

1fF=4log{ε3.70655.0452NRelog[ε1.10982.8257+(7.149NRe)0.8981]}=12.3255fF=0.006583

image

If Fig. 4.2 is used, the chart gives a fM of 0.0265. Thus, the fF is estimated as

fF=0.02654=0.006625

image

Finally, the pressure drop is calculated:

ΔP=ggcρΔz+ρ2gcΔu2=2fFρu2LgcD=32.1732.17(51.57)(966)=51.572(32.17)(0)2=2(0.006625)(51.57)(2.34)2(1000)(32.17)(0.188)=50,435Ibf/ft2=350psi

image

4.3 Single-Phase Gas Flow

The first law of thermodynamics (conservation of energy) governs gas flow in tubing. The effect of kinetic energy change is negligible because the variation in tubing diameter is insignificant in most gas wells. With no shaft work device installed along the tubing string, the first law of thermodynamics yields the following mechanical balance equation:

dPρ+ggcdZ+fMv2dL2gcDi=0 (4.7)

image (4.7)

Because dZ=cos θdL, ρ=29γgPZRTimage, and v=4qsczPscTπDi2TscP,image Eq. (4.7) can be rewritten as

zRT29γgdPP+{ggccosθ+8fMQsc2Psc2π2gcDi5Tsc2[zTP]2}dL=0, (4.8)

image (4.8)

which is an ordinary differential equation governing gas flow in tubing. Although the temperature T can be approximately expressed as a linear function of length L through geothermal gradient, the compressibility factor z is a function of pressure P and temperature T. This makes it difficult to solve the equation analytically. Fortunately, the pressure P at length L is not a strong function of temperature and compressibility factor. Approximate solutions to Eq. (4.8) have been sought and used in the natural gas industry.

4.3.1 Average Temperature and Compressibility Factor Method

If single average values of temperature and compressibility factor over the entire tubing length can be assumed, Eq. (4.8) becomes

z¯RT¯29γgdPP+{ggccosθ+8fMQcs2Psc2z¯2T¯2π2gci5Tsc2P2}dL=0. (4.9)

image (4.9)

By separation of variables, Eq. (4.9) can be integrated over the full length of tubing to yield

Pwf2=Exp(s)Phf2+8fM[Exp(s)1]Qsc2Psc2z¯2T¯2π2gcDi5Tsc2cosθ, (4.10)

image (4.10)

where

s=58γggLcosθgcRz¯T¯. (4.11)

image (4.11)

Eqs. (4.10) and (4.11) take the following forms when U.S. field units (qsc in Mscf/d), are used (Katz et al., 1959):

pwf2=Exp(s)phf2+6.67×104[Exp(s)1]fMqsc2z¯2T¯2di5cosθ (4.12)

image (4.12)

and

s=0.0375γgLcosθz¯T¯ (4.13)

image (4.13)

The Darcy–Wiesbach (Moody) friction factor fM can be found in the conventional manner for a given tubing diameter, wall roughness, and Reynolds number. However, if one assumes fully turbulent flow, which is the case for most gas wells, then a simple empirical relation may be used for typical tubing strings (Katz and Lee, 1990):

fM=0.01750di0.224fordi4.277in. (4.14)

image (4.14)

fM=0.01603di0.164fordi>4.277in. (4.15)

image (4.15)

Guo and Ghalambor (2002) used the following Nikuradse friction factor correlation for fully turbulent flow in rough pipes:

fM=[11.742log(2εdi)]2 (4.16)

image (4.16)

Because the average compressibility factor is a function of pressure itself, a numerical technique such as Newton–Raphson iteration is required to solve Eq. (4.12) for bottom-hole pressure. This computation can be performed automatically with the spreadsheet program Average TZ.xls. Users need to input parameter values in the Input data section and run Macro Solution to get results.

Example Problem 4.2 Suppose that a vertical well produces 2 MMscf/d of 0.71 gas-specific gravity gas through a 278image in. tubing set to the top of a gas reservoir at a depth of 10,000 ft. At tubing head, the pressure is 800 psia and the temperature is 150°F; the bottom-hole temperature is 200°F. The relative roughness of tubing is about 0.0006. Calculate the pressure profile along the tubing length and plot the results.

Solution Example Problem 4.2 is solved with the spreadsheet program AverageTZ.xls. Table 4.1 shows the appearance of the spreadsheet for the Input data and Result sections. The calculated pressure profile is plotted in Fig. 4.3.

Table 4.1

Spreadsheet AverageTZ.xls: The Input Data and Result Sections

AverageTZ.xls
Description: This spreadsheet calculates tubing pressure traverse for gas wells.
Instructions: (1) Input your data in the Input data section; (2) Click “Solution” button to get results; and (3) View results in table and in graph sheet “Profile”.
Input Data
γg= 0.71   
d= 2.259 in.   
ε/d= 0.0006   
L= 10,000 ft   
θ=   
phf= 800 psia   
Thf= 150°F   
Twf= 200°F   
qsc= 2000 Mscf/d   
Solution
fM= 0.017396984   
Depth (ft) T (°R) p (psia) Zav
0 610 800 0.9028
1000 615 827 0.9028
2000 620 854 0.9027
3000 625 881 0.9027
4000 630 909 0.9026
5000 635 937 0.9026
6000 640 965 0.9026
7000 645 994 0.9026
8000 650 1023 0.9027
9000 655 1053 0.9027
10,000 660 1082 0.9028

Image

image
Figure 4.3 Calculated tubing pressure profile for the Example Problem 4.2.

4.3.2 Cullender and Smith Method

Eq. (4.8) can be solved for bottom-hole pressure using a fast numerical algorithm originally developed by Cullender and Smith (Katz et al., 1959). Eq. (4.8) can be rearranged as

PzTdpggccosθ(PzT)2+8fMQsc2Psc2π2gcDi5Tsc2=29γgRdL (4.17)

image (4.17)

that takes an integration form of

PhfPwf[PzTggccosθ(PzT)2+8fMQsc2Psc2π2gcDi5Tsc2]dp=29γgLR. (4.18)

image (4.18)

In U.S. field units (qmsc in MMscf/d), Eq. (4.18) has the following form:

phfpwf[pzT0.001cosθ(pzT)2+0.6666fMqmsc2di5]dp=18.75γgL (4.19)

image (4.19)

If the integrant is denoted with symbol I, that is,

I=pzT0.001cosθ(pzT)2+0.6666fMqsc2di5, (4.20)

image (4.20)

Eq. (4.19) becomes

phfpwfIdp=1875γgL. (4.21)

image (4.21)

In the form of numerical integration, Eq. (4.21) can be expressed as

(pmfphf)(ImfIhf)2+(pwfpmf)(IwfImf)2=18.75γgL, (4.22)

image (4.22)

where pmf is the pressure at the mid-depth. The Ihf, Imf, and Iwf are integrant is evaluated at phf, pmf, and pwf, respectively. Assuming the first and second terms in the right-hand side of Eq. (4.22) each represents half of the integration, that is,

(pmfphf)(ImfIhf)2=1875γgL2 (4.23)

image (4.23)

(pwfpmf)(IwfImf)2=18.75γgL2, (4.24)

image (4.24)

the following expressions are obtained:

pmf=phf+18.75γgLImf+Ihf (4.25)

image (4.25)

pwf=pmf+18.75γgLIwf+Imf (4.26)

image (4.26)

Because Imf is a function of pressure pmf itself, a numerical technique such as Newton–Raphson iteration is required to solve Eq. (4.25) for pmf. Once pmf is computed, pwf can be solved numerically from Eq. (4.26). These computations can be performed automatically with the spreadsheet program Cullender-Smith.xls. Users need to input parameter values in the Input Data section and run Macro Solution to get results.

Example Problem 4.3 Solve the problem in Example Problem 4.2 with the Cullender and Smith Method.

Solution Example Problem 4.3 is solved with the spreadsheet program Cullender-Smith.xls. Table 4.2 shows the appearance of the spreadsheet for the Input data and Result sections. The pressures at depths of 5000 ft and 10,000 ft are 937 psia and 1082 psia, respectively. These results are exactly the same as that given by the Average Temperature and Compressibility Factor Method.

Table 4.2

Spreadsheet Cullender-Smith.xls: The Input Data and Result Sections

Cullender-SmithBHP.xls
Description: This spreadsheet calculates bottom-hole pressure with the Cullender–Smith method.
Instructions: (1) Input your data in the Input data section; and (2) Click Solution button to get results.
Input Data
γg =0.71     
D =2.259 in.     
ε/d =0.0006     
L =10,000 ft     
Θ =0°     
phf =800 psia     
Thf =150°F     
Twf =200°F     
qmsc =2 MMscf/d     
Solution
fM =0.017397     
Depth (ft) T (°R) p (psia) Z p/ZT I
0 610 800 0.9028 1.45263 501.137
5000 635 937 0.9032 1.63324 472.581
10,000 660 1082 0.9057 1.80971 445.349

Image

4.3.3 Flow of Impure Gas

The average temperature average z-factor method and the Cullender and Smith method were derived on the basis of pure gas flow. In reality, almost all gas wells produce certain amount of liquids (oil and water) and sometimes solid particles (sand and coal). The volume fractions of these liquids and solids are low but their effect on pressure can be significant due to their densities being much higher than the density of gas.

A gas-oil-water-sand four-phase flow model was proposed by Guo and Ghalambor (2005) to describe the flow impure gas. The model takes a closed (integrated) form, which makes it easy to use. The Guo–Ghalambor model can be expressed as follows:

b(PPtop)+12bM2ln|(P+M)2+N(Ptop+M)2+N|M+bcNbM2N×[tan1(P+MN)tan1(Ptop+MN)]=aL(cosθ+d2e) (4.27)

image (4.27)

where the group parameters are defined as

a=0.0765γgqg+350γoqo+350γwqw+62.4γsqs4.07Tavqg, (4.28)

image (4.28)

b=5.615qo+5.615qw+qs4.07TavQg, (4.29)

image (4.29)

c=0.00678TavQgA, (4.30)

image (4.30)

d=0.00166A(5.615qo+5.615qw+qs), (4.31)

image (4.31)

e=fM2gDH, (4.32)

image (4.32)

M=cdecosθ+d2e, (4.33)

image (4.33)

N=c2ecosθ(cosθ+d2e)2, (4.34)

image (4.34)

where

A=cross-sectional area of conduit, in.2

DH=hydraulic diameter, ft

fM=Darcy–Wiesbach friction factor (Moody factor)

g=gravitational acceleration, 32.17 ft/s2

L=conduit length, ft

P=pressure, lbf/ft2

Ptop=flowing pressure at section top, lbf/ft2

qg=gas production rate, scf/d

qo=oil production rate, bbl/d

qs=sand production rate, ft3/day

qw=water production rate, bbl/d

Tav=average temperature, °R

γg=specific gravity of gas, air=1

γo=specific gravity of produced oil, freshwater=1

γs=specific gravity of produced solid, fresh water=1

γw=specific gravity of produced water, fresh water=1

The Darcy–Wiesbach friction factor (fM) can be obtained from diagram (Fig. 4.2) or based on fF obtained from Eq. (4.16). The required relation is fM=4fF. The Guo–Ghalambor can also be used for describing mist flow in oil wells (Guo et al., 2008).

Because iterations are required to solve Eq. (4.18) for pressure, a computer spreadsheet program Guo-Ghalambor BHP.xls has been developed.

Example Problem 4.4 For the following data, estimate bottom-hole pressure with the Guo–Ghalambor method:

Total measured depth: 7000 ft
The average inclination angle: 20°
Tubing inner diameter: 1.995 in.
Gas production rate: 1 MMscfd
Gas-specific gravity: 0.7 air=1
Oil production rate: 1000 stb/d
Oil-specific gravity: 0.85 H2O=1
Water production rate: 300 bbl/d
Water-specific gravity: 1.05 H2O=1
Solid production rate: 1 ft3/d
Solid-specific gravity: 2.65 H2O=1
Tubing head temperature: 100°F
Bottom-hole temperature: 224°F
Tubing head pressure: 300 psia

Solution This example problem is solved with the spreadsheet program Guo-GhalamborBHP.xls. The result is shown in Table 4.2. and Table 4.3.

Table 4.3

Result Given by Guo-GhalamborBHP.xls for Example Problem 4.3

Guo-GhalamborBHP.xls
Description: This spreadsheet calculates flowing bottom-hole pressure based on tubing head pressure and tubing flow performance using the Guo–Ghalambor method.
Instruction: (1) Select a unit system; (2) update parameter values in the Input data section; (3) click “Solution” button; and (4) view result in the Solution section.
Input Data U.S. Field Units SI Units
Total measured depth: 7000 ft  
Average inclination angle: 20°  
Tubing inside diameter: 1.995 in.  
Gas production rate: 1,000,000 scfd  
Gas-specific gravity: 0.7 air=1  
Oil production rate: 1000 stb/d  
Oil-specific gravity: 0.85 H2O=1  
Water production rate: 300 bbl/d  
Water-specific gravity: 1.05 H2O=1  
Solid production rate: 1 ft3/d  
Solid-specific gravity: 2.65 H2O=1  
Tubing head temperature: 100°F  
Bottom-hole temperature: 224°F  
Tubing head pressure: 300 psia  
Solution
A= 3.1243196 in.2  
D= 0.16625 ft  
Tav= 622 °R  
cos(θ)= 0.9397014  
(Dρv)= 40.908853  
fM= 0.0415505  
a= 0.0001713  
b= 2.884E–06  
c= 1349785.1  
d= 3.8942921  
e= 0.0041337  
M= 20447.044  
N= 6.669E+09  
Bottom-hole pressure, pwf= 1682 psia  

Image

4.4 Multiphase Flow in Oil Wells

In addition to oil, almost all oil wells produce a certain amount of water, gas, and sometimes sand. These wells are called multiphase oil wells. The TPR equation for single-phase flow is not valid for multiphase oil wells. To analyze TPR of multiphase oil wells rigorously, a multiphase flow model is required.

Multiphase flow is much more complicated than single-phase flow because of the variation of flow regime (or flow pattern). Fluid distribution changes greatly in different flow regimes, which significantly affects pressure gradient in the tubing.

4.4.1 Flow Regimes

As shown in Fig. 4.4, at least four flow regimes have been identified in gas-liquid two-phase flow. They are bubble, slug, churn, and annular flow. These flow regimes occur as a progression with increasing gas flow rate for a given liquid flow rate. In bubble flow, gas phase is dispersed in the form of small bubbles in a continuous liquid phase. In slug flow, gas bubbles coalesce into larger bubbles that eventually fill the entire pipe cross-section. Between the large bubbles are slugs of liquid that contain smaller bubbles of entrained gas. In churn flow, the larger gas bubbles become unstable and collapse, resulting in a highly turbulent flow pattern with both phases dispersed. In annular flow, gas becomes the continuous phase, with liquid flowing in an annulus, coating the surface of the pipe and with droplets entrained in the gas phase.

image
Figure 4.4 Flow regimes in gas-liquid flow. After Goier, G.W., Aziz, K., 1977. The Flow of Complex Mixtures in Pipes. Robert E. Drieger Publishing Co, Huntington, NY.

4.4.2 Liquid Holdup

In multiphase flow, the amount of the pipe occupied by a phase is often different from its proportion of the total volumetric flow rate. This is due to density difference between phases. The density difference causes dense phase to slip down in an upward flow (i.e., the lighter phase moves faster than the denser phase). Because of this, the in-situ volume fraction of the denser phase will be greater than the input volume fraction of the denser phase (i.e., the denser phase is “held up” in the pipe relative to the lighter phase). Thus, liquid “holdup” is defined as

yL=VLV, (4.35)

image (4.35)

where

yL=liquid holdup, fraction

VL=volume of liquid phase in the pipe segment, ft3

V=volume of the pipe segment, ft3

Liquid holdup depends on flow regime, fluid properties, and pipe size and configuration. Its value can be quantitatively determined only through experimental measurements.

4.4.3 TPR Models

Numerous TPR models have been developed for analyzing multiphase flow in vertical pipes. Brown (1977) presents a thorough review of these models. TPR models for multiphase flow wells fall into two categories: (1) homogeneous-flow models and (2) separated-flow models. Homogeneous models treat multiphase as a homogeneous mixture and do not consider the effects of liquid holdup (no-slip assumption). Therefore, these models are less accurate and are usually calibrated with local operating conditions in field applications. The major advantage of these models comes from their mechanistic nature. They can handle gas-oil-water three-phase and gas-oil-water-sand four-phase systems. It is easy to code these mechanistic models in computer programs.

Separated-flow models are more realistic than the homogeneous-flow models. They are usually given in the form of empirical correlations. The effects of liquid holdup (slip) and flow regime are considered. The major disadvantage of the separated-flow models is that it is difficult to code them in computer programs because most correlations are presented in graphic form.

4.4.3.1 Homogeneous-flow models

Numerous homogeneous-flow models have been developed for analyzing the TPR of multiphase wells since the pioneering works of Poettmann and Carpenter (1952). Poettmann–Carpenter’s model uses empirical two-phase friction factor for friction pressure loss calculations without considering the effect of liquid viscosity. The effect of liquid viscosity was considered by later researchers including Cicchitti (1960) and Dukler et al. (1964). A comprehensive review of these models was given by Hasan and Kabir (2002). Guo and Ghalambor (2005) presented work addressing gas-oil-water-sand four-phase flow.

Assuming no slip of liquid phase, Poettmann and Carpenter (1952) presented a simplified gas-oil-water three-phase flow model to compute pressure losses in wellbores by estimating mixture density and friction factor. According to Poettmann and Carpenter, the following equation can be used to calculate pressure traverse in a vertical tubing when the acceleration term is neglected:

Δp=(ρ¯+k¯/ρ¯)Δh144 (4.36)

image (4.36)

where

Δp=pressure increment, psi

ρ¯image=average mixture density (specific weight), lb/ft3

Δh=depth increment, ft

and

k¯=f2Fqo2M27.4137×1010D5 (4.37)

image (4.37)

where

f2F=Fanning friction factor for two-phase flow

qo=oil production rate, stb/day

M=total mass associated with 1 stb of oil

D=tubing inner diameter, ft

The average mixture density ρ¯image can be calculated by

ρ¯=ρ1+ρ22 (4.38)

image (4.38)

where

ρ1=mixture density at top of tubing segment, lb/ft3

ρ2=mixture density at bottom of segment, lb/ft3

The mixture density at a given point can be calculated based on mass flow rate and volume flow rate:

ρ=MVm (4.39)

image (4.39)

where

M=350.17(γ0+WORγw)+GORρairγg (4.40)

image (4.40)

Vm=5.615(Bo+WORBw)+(GORRs)(14.7p)(T520)(z1.0) (4.41)

image (4.41)

and where

γo=oil-specific gravity, 1 for freshwater

WOR=producing water–oil ratio, bbl/stb

γw=water-specific gravity, 1 for freshwater

GOR=producing gas–oil ratio, scf/stb

ρair=density of air, lbm/ft3

γg=gas-specific gravity, 1 for air

Vm=volume of mixture associated with 1 stb of oil, ft3

Bo=formation volume factor of oil, rb/stb

Bw=formation volume factor of water, rb/bbl

Rs=solution gas–oil ratio, scf/stb

p=in-situ pressure, psia

T=in situ temperature, °R

z=gas compressibility factor at p and T.

If data from direct measurements are not available, solution gas–oil ratio and formation volume factor of oil can be estimated using the following correlations:

Rs=γg[p18100.0125API100.00091t]1.2048 (4.42)

image (4.42)

Bo=0.9759+0.00012[Rs(γgγo)0.5+1.25t]1.2 (4.43)

image (4.43)

where t is in-situ temperature in °F. The two-phase friction factor f2F can be estimated from a chart recommended by Poettmann and Carpenter (1952). For easy coding in computer programs, Guo and Ghalambor (2002) developed the following correlation to represent the chart:

f2F=101.4442.5log(Dρv), (4.44)

image (4.44)

where (Dρv) is the numerator of Reynolds number representing inertial force and can be formulated as

(Dρv)=1.4737×105MqoD. (4.45)

image (4.45)

Because the Poettmann–Carpenter model takes a finite-difference form, this model is accurate for only short-depth incremental Δh. For deep wells, this model should be used in a piecewise manner to get accurate results (i.e., the tubing string should be “broken” into small segments and the model is applied to each segment).

Because iterations are required to solve Eq. (4.36) for pressure, a computer spreadsheet program Poettmann-CarpenterBHP.xls has been developed. The program is available from the attached CD.

Example Problem 4.5 For the following given data, calculate bottom-hole pressure:

Tubing head pressure: 500 psia
Tubing head temperature: 100°F
Tubing inner diameter: 1.66 in.
Tubing shoe depth (near bottom hole): 5000 ft
Bottom-hole temperature: 150°F
Liquid production rate: 2000 stb/day
Water cut: 25%
Producing GLR: 1000 scf/stb
Oil gravity: 30 °API
Water-specific gravity: 1.05 1 for freshwater
Gas-specific gravity: 0.65 1 for air

Solution This problem can be solved using the computer program Poettmann-CarpenterBHP.xls. The result is shown in Table 4.4.

Table 4.4

Result Given by Poettmann-CarpenterBHP.xls for Example Problem 4.2

Poettmann–CarpenterBHP.xls
Description: This spreadsheet calculates flowing bottom-hole pressure based on tubing head pressure and tubing flow performance using the Poettmann–Carpenter method.
Instruction: (1) Select a unit system; (2) update parameter values in the Input data section; (3) Click “Solution” button; and (4) view result in the Solution section.
Input Data U.S. Field Units
Tubing ID: 1.66 in
Wellhead pressure: 500 psia
Liquid production rate: 2000 stb/d
Producing gas–liquid ratio (GLR): 1000 scf/stb
Water cut (WC): 25%
Oil gravity: 30 °API
Water-specific gravity: 1.05 freshwater=1
Gas-specific gravity: 0.65 1 for air
N2 content in gas: 0 mole fraction
CO2 content in gas: 0 mole fraction
H2S content in gas: 0 mole fraction
Formation volume factor for water: 1.0 rb/stb
Wellhead temperature: 100°F
Tubing shoe depth: 5000 ft
Bottom-hole temperature: 150°F
Solution
Oil-specific gravity =0.88 freshwater=1
Mass associated with 1 stb of oil =495.66 lb
Solution gas ratio at wellhead =78.42 scf/stb
Oil formation volume factor at wellhead =1.04 rb/stb
Volume associated with 1 stb oil @ wellhead =45.12 cf
Fluid density at wellhead =10.99 lb/cf
Solution gas–oil ratio at bottom hole =301.79 scf/stb
Oil formation volume factor at bottom hole =1.16 rb/stb
Volume associated with 1 stb oil @ bottom hole =17.66 cf
Fluid density at bottom hole =28.07 lb/cf
The average fluid density =19.53 lb/cf
Inertial force (Dρv) =79.21 lb/day-ft
Friction factor =0.002
Friction term =293.12 (lb/cf)2
Error in depth =0.00 ft
Bottom-hole pressure =1699 psia

Image

4.4.3.2 Separated-flow models

A number of separated-flow models are available for TPR calculations. Among many others are the Lockhart and Martinelli correlation (1949), the Duns and Ros correlation (1963), and the Hagedorn and Brown method (1965). Based on comprehensive comparisons of these models, Ansari et al. (1994) and Hasan and Kabir (2002) recommended the Hagedorn–Brown method with modifications for near-vertical flow.

The modified Hagedorn–Brown (mH-B) method is an empirical correlation developed on the basis of the original work of Hagedorn and Brown (1965). The modifications include using the no-slip liquid holdup when the original correlation predicts a liquid holdup value less than the no-slip holdup and using the Griffith correlation (Griffith and Wallis, 1961) for the bubble flow regime.

The original Hagedorn–Brown correlation takes the following form:

dPdz=ggcρ¯+2fFρ¯um2gcD+ρ¯Δ(um2)2gcΔz, (4.46)

image (4.46)

which can be expressed in U.S. field units as

144dpdz=ρ¯+fFMt27.413×1010D5ρ¯+ρ¯Δ(um2)2gcΔz, (4.47)

image (4.47)

where

Mt=total mass flow rate, lbm/d

ρ¯image=in-situ average density, lbm/ft3

um=mixture velocity, ft/s

and

ρ¯=yLρL+(1yL)ρG, (4.48)

image (4.48)

um=uSL+uSG, (4.49)

image (4.49)

where

ρL=liquid density, lbm/ft3

ρG=in-situ gas density, lbm/ft3

uSL=superficial velocity of liquid phase, ft/s

uSG=superficial velocity of gas phase, ft/s

The superficial velocity of a given phase is defined as the volumetric flow rate of the phase divided by the pipe cross-sectional area for flow. The third term in the right-hand side of Eq. (4.47) represents pressure change due to kinetic energy change, which is in most instances negligible for oil wells.

Obviously, determination of the value of liquid holdup yL is essential for pressure calculations. The mH-B correlation uses liquid holdup from three charts using the following dimensionless numbers:

Liquid velocity number, NvL:

NvL=1.938uSLρLσ4 (4.50)

image (4.50)

Gas velocity number, NvG:

NvG=1.938uSGρLσ4 (4.51)

image (4.51)

Pipe diameter number, ND:

ND=120.872DρLσ (4.52)

image (4.52)

Liquid viscosity number, NL:

NL=0.15726μL1ρLσ34, (4.53)

image (4.53)

where

D=conduit inner diameter, ft

σ=liquid–gas interfacial tension, dyne/cm

μL=liquid viscosity, cp

μG=gas viscosity, cp

The first chart is used for determining parameter (CNL) based on NL. We have found that this chart can be replaced by the following correlation with acceptable accuracy:

(CNL)=10Y, (4.54)

image (4.54)

where

Y=2.69851+0.15841X10.55100X12+0.54785X130.12195X14 (4.55)

image (4.55)

and

X1=log(NL)+3. (4.56)

image (4.56)

Once the value of parameter (CNL) is determined, it is used for calculating the value of the group NvLp0.1(CNL)NvG0.575pa0.1NDimage, where p is the absolute pressure at the location where pressure gradient is to be calculated, and pa is atmospheric pressure. The value of this group is then used as an entry in the second chart to determine parameter (yL). We have found that the second chart can be represented by the following correlation with good accuracy:

yLψ=0.10307+0.61777[log(X2)+6]0.63295[log(X2)+6]2+0.29598[log(X2)+6]30.0401[log(X2)+6]4, (4.57)

image (4.57)

where

X2=NvLp0.1(CNL)NvG0.575pa0.1ND. (4.58)

image (4.58)

According to Hagedorn and Brown (1965), the value of parameter ψ can be determined from the third chart using a value of group NvGNL0.38ND2.14image.

We have found that for NvGNL0.38ND2.14>0.01image the third chart can be replaced by the following correlation with acceptable accuracy:

ψ=0.911634.82176X3+1,232.25X3222,253.6X33+116174.3X34, (4.59)

image (4.59)

where

X3=NvGNL0.38ND2.14. (4.60)

image (4.60)

However, ψ=1.0 should be used for NvGNL0.38ND2.140.01image.

Finally, the liquid holdup can be calculated by

yL=ψ(yLψ). (4.61)

image (4.61)

The Reynolds number for multiphase flow can be calculated by

NRe=2.2×102mtDμLyLμG(1yL), (4.62)

image (4.62)

where mt is mass flow rate. The modified mH-B method uses the Griffith correlation for the bubble-flow regime. The bubble-flow regime has been observed to exist when

λG<LB, (4.63)

image (4.63)

where

λG=usGum (4.64)

image (4.64)

and

LB=1.0710.2218(um2D), (4.65)

image (4.65)

which is valid for LB≥0.13. When the LB value given by Eq. (4.65) is less than 0.13, LB=0. 13 should be used.

Neglecting the kinetic energy pressure drop term, the Griffith correlation in U.S. field units can be expressed as

144dpdz=ρ¯+fFmL27.413×1010D5ρLyL2, (4.66)

image (4.66)

where mL is mass flow rate of liquid only. The liquid holdup in Griffith correlation is given by the following expression:

yL=112[1+umus(1+umus)24usGus], (4.67)

image (4.67)

where μs=0.8 ft/s. The Reynolds number used to obtain the friction factor is based on the in-situ average liquid velocity, that is,

NRe=2.2×102mLDμL. (4.68)

image (4.68)

To speed up calculations, the Hagedorn–Brown correlation has been coded in the spreadsheet program Hagedorn Brown Correlation.xls.

Example Problem 4.6 For the data given below, calculate and plot pressure traverse in the tubing string:

Tubing shoe depth: 9700 ft
Tubing inner diameter: 1.995 in.
Oil gravity: 40 °API
Oil viscosity: 5 cp
Production GLR: 75 scf/bbl
Gas-specific gravity: 0.7 air=1
Flowing tubing head pressure: 100 psia
Flowing tubing head temperature: 80°F
Flowing temperature at tubing shoe: 180°F
Liquid production rate: 758 stb/day
Water cut: 10%
Interfacial tension: 30 dynes/cm
Specific gravity of water: 1.05 H2O=1

Solution This example problem is solved with the spreadsheet program HagedornBrownCorrelation.xls. The result is shown in Table 4.5 and Fig. 4.5.

Table 4.5

Result Given by HagedornBrownCorrelation.xls for Example Problem 4.6

HagedornBrownCorrelation.xls
Description: This spreadsheet calculates flowing pressures in tubing string based on tubing head pressure using the Hagedorn–Brown correlation.
Instruction: (1) Select a unit system; (2) update parameter values in the Input data section; (3) click “Solution” button; and (4) view result in the Solution section and charts.
Input Data U.S. Field Units SI Units  
Depth (D): 9700 ft   
Tubing inner diameter (dti): 1.995 in.   
Oil gravity (API): 40 °API   
Oil viscosity (μo): 5 cp   
Production GLR (GLR): 75 scf/bbl   
Gas-specific gravity (γg): 0.7 air=1   
Flowing tubing head pressure (phf): 100 psia   
Flowing tubing head temperature (thf): 80°F   
Flowing temperature at tubing shoe (twf): 180°F   
Liquid production rate (qL): 758 stb/day   
Water cut (WC): 10%   
Interfacial tension (σ): 30 dynes/cm   
Specific gravity of water (γw): 1.05 H2O=1   
Solution
Depth  Pressure  
(ft) (m) (psia) (MPa)
0 0 100 0.68
334 102 183 1.24
669 204 269 1.83
1003 306 358 2.43
1338 408 449 3.06
1672 510 543 3.69
2007 612 638 4.34
2341 714 736 5.01
2676 816 835 5.68
3010 918 936 6.37
3345 1020 1038 7.06
3679 1122 1141 7.76
4014 1224 1246 8.48
4348 1326 1352 9.20
4683 1428 1459 9.93
5017 1530 1567 10.66
5352 1632 1676 11.40
5686 1734 1786 12.15
6021 1836 1897 12.90
6355 1938 2008 13.66
6690 2040 2121 14.43
7024 2142 2234 15.19
7359 2243 2347 15.97
7693 2345 2461 16.74
8028 2447 2576 17.52
8362 2549 2691 18.31
8697 2651 2807 19.10
9031 2753 2923 19.89
9366 2855 3040 20.68
9700 2957 3157 21.48

Image

image
Figure 4.5 Pressure traverse given by Hagedorn Brown Correltion.xls for Example Problem 4.6.

4.5 Summary

This chapter presented and illustrated different mathematical models for describing wellbore/tubing performance. Among many models, the mH-B model has been found to give results with good accuracy. The industry practice is to conduct a flow gradient (FG) survey to measure the flowing pressures along the tubing string. The FG data are then employed to validate one of the models and tune the model if necessary before the model is used on a large scale.

References

1. Ansari AM, Sylvester ND, Sarica C, Shoham O, Brill JP. A comprehensive mechanistic model for upward two-phase flow in wellbores SPE Production and Facilities (May 1994). Trans AIME 1994;297.

2. Brown KE. The Technology of Artificial Lift Methods. Vol. 1 Tulsa, OK: Penn Well Books; 1977;104–158.

3. Chen NH. An explicit equation for friction factor in pipe. Ind Eng Chem Fund. 1979;18:296.

4. Cicchitti A. Two-phase cooling experiments—pressure drop, heat transfer and burnout measurements. Energia Nucleare. 1960;7(6):407.

5. Dukler AE, Wicks M, Cleveland RG. Frictional pressure drop in two-phase flow: a comparison of existing correlations for pressure loss and hold-up. AIChE J 1964;38–42.

6. Duns, H., Ros, N.C.J. Vertical flow of gas and liquid mixtures in wells. In: Proceedings of the 6th World Petroleum Congress, Tokyo, 1963.

7. Goier GW, Aziz K. The Flow of Complex Mixtures in Pipes Huntington, NY: Robert E. Drieger Publishing Co; 1977.

8. Gregory GA, Fogarasi M. Alternate to standard friction factor equation. Oil Gas J. 1985;83:120–127.

9. Griffith P, Wallis GB. Two-phase slug flow. Trans ASME. 1961;83(Ser. C):307–320.

10. Guo B, Ghalambor A. Gas Volume Requirements for Underbalanced Drilling Deviated Holes Tulsa, OK: PennWell Corporation; 2002;132–133.

11. Guo B, Ghalambor A. Natural Gas Engineering Handbook Houston, TX: Gulf Publishing Company; 2005;59–61.

12. Guo B, Sun K, Ghalambor A. Well Productivity Handbook Houston, TX: Gulf Publishing Company; 2008.

13. Hagedorn AR, Brown KE. Experimental study of pressure gradients occurring during continuous twophase flow in small-diameter conduits. J Petroleum Technol 1965;475.

14. Hasan AR, Kabir CS. Fluid Flow and Heat Transfer in Wellbores Richardson, TX: Society of Petroleum Engineers; 2002;10–15.

15. Katz DL, Cornell D, Kobayashi R, Poettmann FH, Vary JA, Elenbaas JR. Handbook of Natural Gas Engineering New York: McGraw-Hill Publishing Company; 1959.

16. Katz DL, Lee RL. Natural Gas Engineering—Production and Storage New York: McGraw-Hill Publishing Company; 1990.

17. Lockhart RW, Martinelli RC. Proposed correlation of data for isothermal two-phase, twocomponent flow in pipes. Chem Eng Prog. 1949;45:39.

18. Moody LF. Friction factor for pipe flow. Trans ASME. 1944;66:671–685.

19. Poettmann FH, Carpenter PG. The multiphase flow of gas, oil, and water through vertical string. Proceedings of the API Drilling and Production Practice Conference held in Dallas, TX 1952;257–263.

Problems

4.1. Suppose that 1000 bbl/day of 16 °API, 5-cp oil is being produced through 2⅞-in., 8.6-lbm/ft tubing in a well that is 3° from vertical. If the tubing wall relative roughness is 0.001, assuming no free gas in tubing string, calculate the pressure drop over 1000 ft of tubing.

4.2. For the following given data, calculate bottom-hole pressure using the Poettmann–Carpenter method:

Tubing head pressure: 300 psia
Tubing head temperature: 100°F
Tubing inner diameter: 1.66 in.
Tubing shoe depth (near bottom hole): 8000 ft
Bottom-hole temperature: 170°F
Liquid production rate: 2000 stb/day
Water cut: 30%
Producing GLR: 800 scf/stb
Oil gravity: 40 °API
Water-specific gravity: 1.05 1 for freshwater
Gas-specific gravity: 0.70 1 for air

4.3. For the data given below, estimate bottom-hole pressure with the Guo–Ghalambor method.

Total measured depth: 8000 ft
The average inclination angle:
Tubing inner diameter: 1.995 in.
Gas production rate: 0.5 MMscfd
Gas-specific gravity: 0.75 air=1
Oil production rate: 2000 stb/d
Oil-specific gravity: 0.85 H2O=1
Water production rate: 500 bbl/d
Water-specific gravity: 1.05 H2O=1
Solid production rate: 4 ft3/d
Solid-specific gravity: 2.65 H2O=1
Tubing head temperature: 100°F
Bottom-hole temperature: 170°F
Tubing head pressure: 500 psia
Tubing shoe depth: 6000 ft
Tubing inner diameter: 1.995 in.
Oil gravity: 30 °API
Oil viscosity: 2 cp
Production GLR: 500 scf/bbl
Gas-specific gravity: 0.65 air=1
Flowing tubing head pressure: 100 psia
Flowing tubing head temperature: 80°F
Flowing temperature at tubing shoe: 140°F
Liquid production rate: 1500 stb/day
Water cut: 20%
Interfacial tension: 30 dynes/cm
Specific gravity of water: 1.05 H2O=1

4.4. For the data given below, calculate and plot pressure traverse in the tubing string using the Hagedorn–Brown correlation:

4.5. Suppose 3 MMscf/d of 0.75 specific gravity gas is produced through a 31/2-in. tubing string set to the top of a gas reservoir at a depth of 8000 ft. At the tubing head, the pressure is 1000 psia and the temperature is 120°F; the bottom-hole temperature is 180°F. The relative roughness of tubing is about 0.0006. Calculate the flowing bottom-hole pressure with three methods: (1) the average temperature and compressibility factor method; (2) the Cullender–Smith method; and (3) the four-phase flow method. Make comments on your results.

4.6. Solve Problem 4.5 for gas production through a K-55, 17-lb/ft, 51/2-in casing.

4.7. Suppose 2 MMscf/d of 0.65 specific gravity gas is produced through a 27/8-in. (2.259-in. inside diameter) tubing string set to the top of a gas reservoir at a depth of 5000 ft. Tubing head pressure is 300 psia and the temperature is 100°F; the bottom-hole temperature is 150°F. The relative roughness of tubing is about 0.0006. Calculate the flowing bottom pressure with the average temperature and compressibility factor method.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
18.227.46.229