Anda di halaman 1dari 13

J Petrol Explor Prod Technol (2018) 8:799–811

https://doi.org/10.1007/s13202-017-0397-0

ORIGINAL PAPER - PRODUCTION ENGINEERING

Transient flowing-fluid temperature modeling in reservoirs


with large drawdowns
N. Chevarunotai1 • A. R. Hasan2 • C. S. Kabir3 • R. Islam2

Received: 30 April 2017 / Accepted: 1 October 2017 / Published online: 14 November 2017
Ó The Author(s) 2017. This article is an open access publication

Abstract Modern downhole temperature measurements Keywords Transient heat transport in porous media 
indicate that bottomhole fluid temperature can be signifi- Joule–Thomson effect  Heat transport to under- and
cantly higher or lower than the original reservoir temper- overburden formations  Analytical solutions  Validation
ature, especially in reservoirs where high-pressure of analytical solutions with numerical results
drawdown is expected during production. This recent
finding contradicts the isothermal assumption originally
Abbreviation
made for routine calculations. In a high-pressure drawdown
A Flow area, ft2, L2
environment, the Joule–Thomson (J–T) phenomenon plays
Bo Oil formation volume factor, bbl/STB
an important role in fluid temperature alteration in the
cp System specific heat capacity, Btu/lbm °F, L2/t2T
reservoir. This paper presents a robust analytical model to
cpf Formation specific heat capacity, Btu/lbm °F, L2/t2T
estimate the flowing-fluid temperature distribution in a
cpo Oil specific heat capacity, Btu/lbm °F, L2/t2T
reservoir that accounts for the J–T heating or cooling
cpw Water specific heat capacity, Btu/lbm °F, L2/t2T
effect. All significant heat transfer mechanisms for fluid
CJ Joule–Thomson coefficient, °F/psi, TLt2/m
flow in the reservoir, including heat transfer due to con-
h Formation thickness, ft, L
vection, J–T phenomenon, and heat transfer from over-
hc Heat transfer coefficient, Btu/hr ft2 °F, m/t3/T
burden and under-burden formations, are incorporated in
H^ Enthalpy, lbm ft2/hr2, mL2/t2
this study. The proposed model successfully validates the
results of a rigorous numerical model that intrinsically k Reservoir permeability, md, L2
honored field data. p Pressure, psi, m/Lt2
pb Bubble point pressure, psi, m/Lt2
pe Pressure at reservoir external boundary, psi, m/Lt2
& C. S. Kabir pi Initial reservoir pressure, psi, m/Lt2
ckabir@Central.UH.EDU; shahkabir@gmail.com pwf Flowing-fluid pressure at well bottom, psi, m/Lt2
N. Chevarunotai Pe Peclet number (= ur/a), dimensionless
note_natash@hotmail.com q Volumetric flow rate, ft3/hr, L3/t
A. R. Hasan
*
q Conductive heat transport, Btu/hr ft2, m/Lt3
rashid.hasan@tamu.edu Q_ Net heat transfer rate between the system and
R. Islam surroundings, Btu/hr ft2, m/Lt3
rislam@tamu.edu r Radius, ft, L
1 re External reservoir radius, ft, L
Chevron Thailand Exploration and Production Ltd., Bangkok,
Thailand rw Wellbore radius, ft, L
2 S Saturation
Department of Petroleum Engineering, Texas A&M
University, College Station, TX 77843, USA So Oil saturation
3 Sw Water saturation
Department of Petroleum Engineering, University of
Houston, 5000 Gulf Freeway, Houston, TX, USA t Time, hr, t

123
800 J Petrol Explor Prod Technol (2018) 8:799–811

T Fluid temperature, °F, T unaccounted for a given low intrinsic flow rate in heavy-oil
Te Fluid temperature at reservoir external boundary, °F, reservoirs.
T The J–T heating or cooling effect originated from
Ti Initial reservoir temperature, °F, T interpreting temperature logs. Steffensen and Smith (1973)
Ts Temperature of overburden and under-burden proposed an analytical solution for estimating the fluid’s
formations, °F, T static and flowing temperature at bottomhole during
Twf Flowing-fluid temperature at well bottom, °F, T steady-state flow by incorporating the J–T effect. They
*
u Superficial velocity, ft/hr, L/t pointed out that the main heat transfer mechanisms of
ur Fluid local velocity in radial direction, ft/hr, L/t fluids in the reservoir during production and injection are
^ Fluid internal energy, lbm ft2/hr2, mL2/t2 heat convection and J–T heating (or cooling). Temperature
U
change due to radial conduction is normally negligible.
^
V Specific volume, ft3/lbm, L3/m
They also proposed that heat transfer between reservoir and
k Reservoir thermal conductivity, Btu/hr ft °F, TLt2/m
overburden and under-burden formations during steady-
a Thermal diffusivity (= k/qcp), ft2/hr, L2/t
state flow is negligible; therefore, the ‘‘heat transfer to
l Fluid viscosity, cp, m/Lt
overburden’’ term was not included in their study. Kabir
q Density, lbm/ft3, m/L3
et al. (2012), among others, showed how independent
qo Oil density, lbm/ft3, m/L3
estimation of individual layer contributions may be made
qw Water density, lbm/ft3, m/L3
from temperature profiles in both gas and oil wells, with
qf Formation density, lbm/ft3, m/L3
the J–T effect playing a major role. More recently, Onur
ro Joule–Thomson throttling coefficient of oil, Btu/
and Palabiyik (2015) offered an analytical solution for
lbm psi, L3/m
single-phase water flowing in a geothermal reservoir that
rw Joule–Thomson throttling coefficient of water, Btu/
accounted for the effect of skin. Their approach was to
lbm psi, L3/m
study the use of temperature data for estimating reservoir
*
s Stress, lbf/ft2, m/Lt2
parameters by history matching. Onur and Cinar (2016)
/ Porosity presented an analytical solution accounting for the J–T
effect, but not heat exchange with the overburden and
under-burden formations.
High-pressure drawdown is normally required to com-
Introduction mercially produce from challenging reservoirs, such as
those in deep, low-permeability, and overpressure systems.
Most reservoir engineering calculations presuppose that the As a result, the impact of the J–T effect on flowing-fluid
fluid temperature entering the wellbore has the same tem- temperature is more prominent in some of the deepwater
perature as that in the reservoir, regardless of pressure drop reservoirs in Gulf of Mexico. In some cases, the J–T effect
and elapsed time. While the assumption of constant fluid may trigger fluid temperature increase of 20–30 °F higher
temperature may be true for high-permeability systems, than the fluid temperature at the initial-reservoir condition.
reservoirs undergoing production from low-permeability Yoshioka et al. (2005, 2006) introduced a coupled
systems at significant drawdowns may not conform to this reservoir-and-wellbore analytical temperature model for
simplified assumption. This reality has prompted several horizontal well production in a single-phase reservoir,
studies to probe a radial distribution of the fluid tempera- assuming steady-state conditions. An extended version of
ture in time. Yoshioka et al.’s work was presented by Dawkrajai et al.
Early attempts to establish fluid temperature were (2006). They developed a finite-difference solution for the
mainly for heavy-oil reservoir management in thermal coupled reservoir/wellbore system to estimate fluid tem-
recovery operations. One of the earliest models for esti- perature distribution in the reservoir for two-phase pro-
mating temperature distribution during steam injection was duction in horizontal wells. Their numerical solution
presented by Lauwerier (1955). Subsequently, several removes the steady-state assumption and allows variation
models were presented by Spillette (1965) and Satman of reservoir and fluid properties in space and time.
et al. (1979) using different approaches. More recently, Tan Duru and Horne (2010) developed a semianalytical
et al. (2012) compared some of these solutions and offered solution for the same problem, taking into account J–T
a solution of their own. All of these models treated heat heating or cooling, as well as heat conduction and con-
conduction and convection as main heat transfer mecha- vection. They applied the operator-splitting and time-
nisms in the reservoir; however, fluid temperature change stepping (OSATS) semianalytical technique to solve the
due to the Joule–Thomson (J–T) effect remained problem and split the reservoir energy-balance equation
into two parts: convective transport and diffusion. They

123
J Petrol Explor Prod Technol (2018) 8:799–811 801

solved the convective transport part analytically and used transfer from a reservoir to overburden and under-burden
the solution to approach the diffusion part. The diffusion formations is incorporated in this model formulation fol-
part of the energy-balance equation was solved semiana- lowing App’s approach. The model is validated with results
lytically; that is, the result from the first timestep is the from the rigorous numerical model developed earlier by
initial condition for next timestep, and so on. They also App (2010), based on one set of actual field data. Model
coupled the reservoir temperature model with Izgec et al.’s validation shows that the estimated temperature values
(2007) analytical wellbore temperature model for fluid compare favorably with those obtained from the numerical
temperature analysis in the entire production system. simulator.
App (2009, 2010) developed a nonisothermal reservoir
simulator for single-phase oil flow by coupling mass and
energy-balance equations with Darcy’s law. He included Model development
all possible heat transfer mechanisms in the reservoir as
part of the comprehensive energy-balance equation. While Reservoir system
earlier work by other authors in this area generally assumes
no heat transfer from a reservoir to its surroundings (adi- The reservoir system considered in this study is a 1D radial
abatic process), App’s work incorporates potential heat reservoir where fluid flow occurs only in the radial direc-
transfer from reservoir to overburden and under-burden tion. The only flowing fluid in the reservoir is oil, and there
formations. His model shows that heat loss to overburden is no free gas in the system. Connate water remains
strata is significant and becomes crucial when fluid is immobile. Figure 1 shows a schematic of the reservoir
significantly heated up later in the production period. He system used in the study. We note that fluid flow in the
also discussed potential change in well productivity due to idealized circular reservoir occurs in the ‘‘negative’’ r-di-
J–T heating (or cooling) in high-pressure, high-drawdown rection. Figure 1 displays this simplified wellbore and
reservoirs because fluid viscosity variation depends on reservoir configuration.
temperature change. Ramazanov et al. (2013) proposed a
similar numerical model that included convection, radial Comprehensive energy-balance equation
heat conduction, and the J–T effect. Their numerical model
validated their earlier (2007) study and pointed out that the A principle for estimation of fluid temperature distribution
impact of radial conduction to fluid temperature distribu- in the reservoir is conservation of energy in the system,
tion in a reservoir is minimal when the production rate which includes reservoir fluid and rock. Conservation of
remains constant. mass for reservoir fluids is also incorporated to achieve a
Recently, App and Yoshioka (2013) offered an analyt- comprehensive energy-balance equation of the system. We
ical solution for steady-state fluid temperature change as a also assume that reservoir is perfectly horizontal; thus,
function of producing rate, reservoir permeability, and gravitational effect (change in fluid potential energy) is
drawdown, among other variables. They used Peclet negligible.
number (Pe = ur/a) to combine production rate and for- The general form of thermal energy balance in terms of
mation thermal conductivity to emphasize the effect of Pe equation of change for internal energy (Bird et al. 2006)
on fluid temperature change. They also pointed out that the can be written as:
effect of permeability is included through Pe as it incor-
porates fluid velocity. Their study clearly shows that at
high reservoir thermal conductivity when Pe \ 1, the fluid
temperature change is strongly influenced by Pe due to
rapid conduction of heat through the formation. However,
for Pe [ 3, the influence of Pe on steady-state fluid tem-
perature is negligible. The study also shows that fluid
temperature changes minimally at very low Pe (\ 0.1),
which appears quite reasonable.
This paper presents an analytical transient-temperature
model for estimating the flowing-fluid temperature distri-
bution in a single-phase oil reservoir with constant rate
production. The paper also presents an application of this
approach to single-phase gas reservoirs. The J–T effect is
included as one of the main energy transformation mech-
anisms of fluid flow in the reservoir. Additionally, heat Fig. 1 Schematic of the wellbore/reservoir system configuration

123
802 J Petrol Explor Prod Technol (2018) 8:799–811

o ^         Details of the comprehensive energy-balance equation are


^ *  r  q*  p r  u*  *s : ru*
qU ¼  r  qUu
ot given in ‘‘Appendix A.’’
þ Q_ ð1Þ

where U ^ is fluid internal energy, q is fluid and/or rock Analytical model


*
density, and u is fluid local velocity. The r term generally
We rearranged the comprehensive energy-balance equation
represents the net input rate of energy per unit volume of
of the system by applying all the assumptions described in
the system. The first term on the left side of Eq. (1) rep-
‘‘Appendix A.’’ The energy-balance equation can be
resents the total rate of internal energy increase in the
reduced to a first-order, partial-differential equation (PDE):
system. The first and second terms on the right side are net
input rate of internal energy to the system caused by con- oT B oT C D E
  2¼ Tþ ð3Þ
vective transport and heat conduction, respectively. The ot Ar or Ar A A
third term represents the net reversible rate of internal The method of characteristics was used to solve the PDE
energy increase due to fluid compression (pressure differ- to arrive at a final form of the proposed analytical solution
ence), while the fourth term is the net irreversible rate of given below. ‘‘Appendix B’’ presents the details of this
internal energy increase caused by fluid viscous dissipa- derivation.
tion. The fourth term is also referred to as ‘‘J–T’’ in this  
C HðAr2 þ2BtÞ H ðAr 2 þ 2BtÞ
study. T ðr; tÞ ¼ Ti þ e 2B Ei 
In addition to heat conduction, convection, and the J–T 2B   2B
phenomenon caused by fluid flow in the reservoir, energy C HAr2 HAr 2
 e 2B Ei  ð4Þ
transfer from surroundings (overburden and under-burden 2B 2B
formations) to the system (reservoir fluids and formation) is where
considered in this study. Therefore, a term representing the 
  2ph
net energy transfer rate between the system and sur- A ¼ £so qo cpo þ £sw qw cpw þ ð1  £Þqf cpf ð5Þ
_ is added to the energy-balance equation as q
roundings, Q,
the last term in Eq. (1). B ¼ qo cpo ð6Þ
Using the principles of rock and fluid enthalpy, Fourier’s qqo ro l
law of conduction, Newton’s law of cooling, conservation C¼ ð7Þ
2phk
of mass, and Darcy’s law, Eq. (1) is rearranged and
4hc p
rewritten as D¼ ð8Þ
q
  oT oT
£so qo cpo þ £sw qw cpw þ ð1  £Þqf cpf þ qo ur cpo 4hc p
ot or E¼ Ti ð9Þ
op op q
þ qo ur ro þ ½£so qo ro þ £sw qw rw  1
 or ot and
1o oT
¼ kr þ Q_ H¼
D
ð10Þ
r or or A
ð2Þ
Equation (4) can be used directly with any software that
Equation (2) is considered a comprehensive energy- requires fluid bottomhole temperature as an input, such as
balance equation for the system of interest. The first term in well pressure/temperature traverse calculations and
on the left side of Eq. (2) contains the heat capacity of oil, production logging. When we solve Eq. (3) by assuming
water, and rock, which collectively represent energy fluid property variation to be negligible, greater accuracy in
change due to temperature transient. Similarly, the estimated sandface temperature can be achieved when
second term represents convective heat transport. The radial segments in the reservoir facilitate variation of fluid
third term is energy change due to J–T effect, and the properties from one node to the next. For temperature
fourth term represents energy change due to pressure computation with Eq. (4), we allowed such variation of
transient in the reservoir. This pressure transient term is fluid properties with pressure and temperature using 100
neglected in deriving the analytical solution presented radial segments with logarithmic spacing in the reservoir.
below. The first term on right side reflects change in energy Calculation initiates with known pressure, temperature, and
arising from radial heat conduction, and the last term fluid property values at the reservoir boundary. Analytical
represents rate of heat transfer across system boundary, expressions then facilitate estimation of pressure and
meaning to overburden and under-burden formations. temperature at the next node. During these computations,

123
J Petrol Explor Prod Technol (2018) 8:799–811 803

property values are retained from the prior node. New production at a constant rate. One observes that the tem-
pressure and temperature then allow computation of new peratures estimated with the proposed model are very close
property values. This procedure repeats itself until the to App’s rigorous, numerical solutions.
wellbore is reached. Because pressure, temperature, and Figure 3 compares the analytical solutions for the
viscosity change much faster as the wellbore is approached, sandface oil temperature (solid lines) with those of App
logarithmic grid spacing—with a shorter spatial step at (dashed lines). This figure reveals the evolution of flowing-
wellbore’s proximity—works very well. We have fluid temperature at the sandface with time for the same
investigated the effect of a number of spatial nodes on constant rate production scenarios as before. This fig-
computational accuracy and found 100 nodes to be quite ure instills confidence in that the simplified analytical
satisfactory. model yields very reasonable estimations of the bottomhole
temperature for different production rate scenarios.
Figure 3 shows that for any constant production rate, the
Model validation oil temperature rises rapidly with time and then flattens out.
Indeed, for high production rates, oil temperature actually
App’s (2010) simulated results, which were anchored in begins a slow decline with time after attaining the maxi-
field data, formed the cornerstone for model validation. mum value. Figure 4 presents this phenomenon in a dif-
This approach also implicitly verifies the results of the ferent way. Each profile of different color represents the
proposed analytical model with those of rigorous numerical flowing-fluid temperature distribution in the reservoir at a
solutions offered by App. ‘‘Appendix C’’ presents the particular producing time for a production rate of 6200
reservoir and heat transfer parameters used in these cal- STB/D. The solid lines represent our analytical solution,
culations. In all cases, solid lines represent the reservoir and the dashed lines are from App’s study. Again, we
temperature profiles estimated with the proposed analytical observe that results from the analytical model are very
model. In all subsequent discussions, reference to estimates close to the results obtained from App’s numerical simu-
using the analytical model implies the use of Eq. (4), with lations. Differences in temperature profiles between the
the parameters estimated with Eq. (5) through Eq. (10), analytical model and App’s simulator are expected because
thereby allowing all fluid properties to vary with pressure of the significantly more assumptions made to arrive at the
and temperature. analytical solution.
In this example, the flowing-fluid temperature distribu- Examination of Fig. 4 offers several insights. We
tion in the reservoir is calculated for five different constant observe that oil temperature in most of the formation
production rates: 970, 2050, 3270, 4650, and 6200 STB/D. remains unaffected; temperature increase is only noticeable
Figure 2 presents a comparison of solution generated with up to about 100 ft from the wellbore for long producing
Eq. (4) with those of App. The solid lines represent the times. The explanation of this phenomenon is simple; most
solutions of the analytical model, and the dashed lines do of the pressure drop—the cause for temperature rise—oc-
the same for those of App. Each profile represents distri- curs near the wellbore, especially for shorter production
bution of the flowing-fluid temperature in the reservoir at periods. In addition, because heat is generated continuously
different production rates, after 50 days of continuous due to J–T effect, longer production leads to greater tem-
perature rise. However, the rate of temperature increase
slows down with time, and finally the reversal occurs.
330
Therefore, fluid temperature rise for 400 days (black lines)
6200 STB/D
325
Analycal
---------- Numerical 330
Fluid temperature, oF

320 4650 STB/D


325 6200 STB/D
Fluid temperature, oF

315 3270 STB/D


4650 STB/D
320
2050 STB/D 3270 STB/D
310 315
2050 STB/D
970 STB/D
305 310
970 STB/D
305
300 Analycal ----------- Numerical
1.E-01 1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 300
Radial distance from wellbore,  0 100 200 300 400 500
Time, days
Fig. 2 Temperature estimations of analytical model at different rates
compare favorably with those of App’s (2010) numerical model for Fig. 3 Results of the analytical model compare well with App’s
50 days of production numerical model (2010) for the sandface oil temperature

123
804 J Petrol Explor Prod Technol (2018) 8:799–811

330 ‘‘Appendix C.’’ Viscosity, in turn, influences flowing


100 days pressure gradient, and consequently, the reservoir pressure.
325
400 days However, fluid temperature depends on heat generation due
10 days
to fluid expansion, which depends on the pressure gradient,
Fluid temperature, oF

Analycal --------Numerical
320 1 day
dp/dr. Figure 4 shows this complex interdependency of
315 3 hr fluid viscosity, pressure, and temperature in temperature
trends.
310
With the increase in production time, increasing fluid
temperature causes lowering of oil viscosity. For a constant
305
flow rate, lower viscosity causes lower pressure drawdown,
300 resulting in higher reservoir pressure than if viscosity had
1.E-01 1.E+00 1.E+01 1.E+02 1.E+03 1.E+04
remained constant. Higher pressure, however, triggers
Radial distance from wellbore, 
increase in oil viscosity, which contributes to increased
Fig. 4 Analytical model agrees well with App’s (2010) numerical pressure gradient near the wellbore. This increased pres-
model results at different times for the 6200 STB/D producing rate sure gradient in both the analytical and numerical models
causes the fluid temperature (black solid and dashed lines
of production is less than that for 100 days (blue lines). in Fig. 4) to be lower than that after 100 days of production
This reversal is captured in both App’s numerical and our (blue lines) after 400 days of production. To investigate the
analytical solutions. effect of viscosity variation with temperature, we regen-
Let us discuss the two reasons related to reduction and erated the solutions with constant fluid viscosity by keep-
ultimate reversal in temperature rise with time. The pri- ing all other input parameters the same as that for the
mary reason is that our model accounts for fluid heat loss to rigorous model. Figure 6 presents those results with solid
the overburden and under-burden formations. This heat lines representing rigorous solution (with property varia-
loss increases with increased fluid temperature. Figure 5 tion), while the dashed lines are for constant viscosity
shows the estimated fluid temperature using the rigorous condition.
model (solid lines) compared to that estimated assuming no Lower temperatures in Fig. 6 for lower producing rates
heat loss (everything else remaining the same) to the for- represent those with constant viscosity. However, for
mation (dashed lines). Note that the maximum temperature higher producing rates, the constant viscosity model gen-
after 400 days of flow period is about 10 °F higher when erally estimates higher temperatures than does the rigorous
heat loss to the formation is not accounted for compared to (variable viscosity) model. Even for higher producing rates
when it is. Further analyses of ignoring fluid heat loss to the though, the trend reverses with producing times. Again, the
formation are discussed later. interdependence of pressure, temperature, and viscosity
The other reason for reduced temperature rise with precipitate these complex temperature profiles. Given the
producing time is that oil viscosity depends on temperature significant discrepancy in temperature profiles exhibited in
and pressure. We have used viscosity data from laboratory Fig. 6, we use the temperature-dependent viscosity as
measurements for this particular reservoir fluid as pre- default.
sented by App (2010) and is reproduced in Fig. 8 in

Fig. 5 Comparison of 335


analytical models without heat
transfer and with heat transfer 400 days
330
and viscosity variation --------- w/o heat transfer w/ heat transfer
100 days
Fluid temperature, oF

325

320 10 days

315
1 day

3 hr
310

305

300
1.E-01 1.E+00 1.E+01 1.E+02 1.E+03 1.E+04
Radial distance from wellbore, 

123
J Petrol Explor Prod Technol (2018) 8:799–811 805

Fig. 6 Comparison of results of 330


the analytical models with and
without viscosity variation 6200 STB/D
325

Fluid temperature, oF
Constant viscosity Variable viscosity
4650 STB/D
320
3270 STB/D
315
2050 STB/D
310
970 STB/D
305

300
0 50 100 150 200 250 300 350 400 450
Time, hr

In modeling temperature distribution in the reservoir, the small spatial steps, the gas compressibility may be still
many investigators have neglected heat transfer to/from kept small so that Eq. (4) may be applied to a gas reservoir
overburden and under-burden formations Q. _ This assump- as well. Using this approach, we estimated the sandface gas
tion simplifies the modeling approach and results in a much temperature as a function of producing time and that is
simpler expression for the fluid temperature as a function shown in Fig. 7. The input values, provided in ‘‘Appendix
of radial distance and time; ‘‘Appendix B’’ presents this E,’’ were taken from App’s (2009) study. Good agreement
development. In addition, as Fig. 3 shows, neglecting for- of our estimates with those of App’s numerical solution is
mation heat loss does not appear to cause significant errors evident in Fig. 7.
at low drawdowns. However, for higher flow rates and later
times, the estimation error can be quite large; Chevarunotai
(2014) presents further discussion on this topic. Discussion

The analytical reservoir temperature solution derived for


Model’s application in gas reservoirs the simplified reservoir system was validated with App’s
numerical model. Although we made several assumptions
In arriving at the analytical solution for the fluid temper- in our study to simplify the problem to 1D radial reservoir
ature for fluid flow from the reservoir bulk toward the system, the flowing-fluid temperatures estimated by the
wellbore, we assumed the fluid to be only slightly com- proposed analytical solutions are very comparable to those
pressible, thereby allowing the PDE [Eq. (3)] to be linear. calculated by App’s rigorous numerical solution, as shown
For gas, this assumption of small compressibility is gen- in all case studies. The simple form of the analytical
erally untrue. However, by dividing the reservoir into many solutions allows anyone to adopt and apply it to the
radial segments with logarithmic spacing, thereby keeping problems of interest. The proposed solution can also be

Fig. 7 Bottomhole temperature 4500 375


with time for the low-pressure
4000 - - - App's model 370
gas
This study
3500 365
3000 360
pwf, psia

Twf, oF

2500 355
2000 350
1500 345
1000 340
500 335
0 330
0.0 0.5 1.0 1.5 2.0 2.5 3.0
t, hr

123
806 J Petrol Explor Prod Technol (2018) 8:799–811

applied to more complex problems, such as in reservoirs of thermal conduction can be neglected in nonfractured
other geometry by considering the Dietz (1965) shape geothermal systems because Pe [ 1 is satisfied. For the
factor. range of cases that we investigated, Pe is greater than 5.6,
The impact of Joule–Thomson phenomenon is actually a and the effect of permeability, as well as thermal conduc-
function of pressure drop across the reservoir, fluid flow tivity, is negligible within engineering accuracy. We note
rate, and the J–T throttling coefficient. In general, the J–T that any significant change in the wellbore flowing-fluid
coefficient is positive in low-pressure gas reservoir and is temperature is likely to occur in overpressure reservoirs for
negative in high-pressure gas reservoirs and oil reservoirs sustaining high-flow rates, and, in turn, high Pe. This
of any pressure range. As a result, J–T heating becomes the reality increases the likelihood of application of the pro-
norm in low-to-high-pressure oil and high-pressure posed analytical model.
([ 7000 psi) gas reservoirs, whereas J–T cooling occurs in These model results show that fluid temperature in the
low-pressure (\ 5000 psi) gas reservoirs. The approach to near-wellbore region can be significantly different from the
estimate the flowing-fluid temperature presented in this original reservoir temperature during production. A rea-
paper can be used as a basis and adapted for reservoir sonable estimation of the bottomhole flowing-fluid tem-
temperature estimation in gas reservoirs. perature assists in well design from the standpoints of
Based on model results, we surmise that fluid tempera- equipment selection and management of annular-pressure
ture in the near-wellbore region can be significantly dif- buildup or APB.
ferent from the original reservoir temperature. An accurate
estimation of the reservoir fluid temperature from the
analytical formulations can yield a better estimation of well Conclusions
productivity, which is useful in production optimization
and well development planning. A reasonable estimation of This paper presents an analytical model for the flowing-
the bottomhole flowing-fluid temperature is also advanta- fluid temperature estimation in a single-phase oil reservoir.
geous in well design from the standpoints of equipment Concepts of energy balance and conservation of mass were
selection and management of annular-pressure buildup or applied to arrive at an analytical formulation to evaluate
APB (Oudeman and Kerem 2006; Hasan et al. 2010). fluid temperature in a reservoir producing at a constant
When J–T heating is pronounced, excessive heating of rate. Fluid temperature change due to the Joule–Thomson
annular fluid may result, thereby triggering APB. effect, as well as energy exchange between a reservoir and
The proposed analytical solution also improves wellbore its surroundings (overburden and under-burden forma-
fluid and heat flow modeling because of more realistic tions), was incorporated in this study. Results from a rig-
temperature evaluation at sandface. The bottomhole flow- orous numerical model validated the simplified analytical
ing-fluid temperature derived from the analytical model model within engineering accuracy. Therefore, we reached
can be coupled with wellbore heat transfer model to allow the following conclusions:
prediction of flowing-fluid temperature along the wellbore.
1. The proposed analytical model provides comparable
Accurate flowing-fluid temperature profile along the well-
reservoir temperature estimation to the rigorous
bore is also desirable for well design and production
numerical simulator developed by App (2010), which
optimization, as well as for pressure transient analysis
is anchored in actual field data. Calculations of Peclet
(Onur and Cinar 2016). An accurate estimation of the
numbers suggested that ignoring conductive heat
reservoir fluid temperature from the analytical formulations
transport appears reasonable for field production rates
can yield a better estimation of well productivity index,
of interest within the scope of this investigation.
which is useful in production optimization and well
Generally speaking, an analytical model is relatively
development planning.
simpler and allows the calculations to be performed in
To arrive at the analytical solution, we omitted radial
a spreadsheet.
heat conduction, meaning the influence of Peclet number
2. The advantage of this analytical model over other
on fluid temperature change has been ignored. In ‘‘Ap-
analytical solutions for reservoir temperature estima-
pendix C,’’ we show that for the reservoir and fluid prop-
tion is that heat transfer from/to overburden and under-
erties used in this study, Peclet numbers for all flow rates
are higher than 5.6. As App and Yoshioka (2013) showed, burden formations Q_ is included. While the derivation
the influence of Peclet number on fluid temperature is of the analytical solution neglects property variation,
negligible when it exceeds 3. They also noted that because the use of the solution allows for property changes
formation permeability affects oil production rate, it can with pressure and temperature. We have shown that Q_
influence oil temperature increase due to expansion. Ear- is crucial in the estimation of flowing-fluid temperature
lier, Shook (2001) also reached a similar conclusion in that in a reservoir, especially at long producing times when

123
J Petrol Explor Prod Technol (2018) 8:799–811 807

 
the reservoir fluid is heated significantly, and the o 1o ^  1o oT op Dp
reservoir fluid temperature is very different from that qH þ rqHur ¼ kr  ur þ þ Q_
ot r or r or or or Dt
in its surroundings.
ð13Þ
3. Accounting for viscosity variation with temperature
and pressure enhances the accuracy of temperature Applying mass conservation in 1D radial system,
op
estimation.
otþ 1r oro ðrqur Þ ¼ 0, Eq. (13) can be rewritten as
4. The analytical model can be further extended to gas  
oH 1 o oT op _
reservoirs by accounting for changing properties, such q ¼ kr þ þ Q: ð14Þ
as density, viscosity, and the J–T coefficient with ot r or or ot
pressure and temperature. Enthalpy of the reservoir rock depends only on its
temperature, which is given by dHf = cpfdT. However, the
Open Access This article is distributed under the terms of the enthalpy of a fluid depends both on its temperature and
Creative Commons Attribution 4.0 International License (http:// pressure; that is,
creativecommons.org/licenses/by/4.0/), which permits unrestricted  
use, distribution, and reproduction in any medium, provided you give oH oH
dH ¼ dT þ dp ¼ cp dT  CJ cp dp ð15Þ
appropriate credit to the original author(s) and the source, provide a oT p op T
link to the Creative Commons license, and indicate if changes were
made. In Eq. (15), cp is the oil specific heat and CJ is the Joule–
Thomson coefficient. Therefore, reservoir oil enthalpy is
expressed by Eq. (16):
Appendix A: Comprehensive energy-balance
equation of the system dHo ¼ cpo dT þ ro dp; ð16Þ
where ro represents the product, - CJ cp. A similar
The energy conservation is the underlying principle for the
expression for the connate water can be written.
estimation of fluid temperature distribution in the reservoir,
We write enthalpy in terms of pressure and temperature
involving rock and fluid. Conservation of mass for reser-
for each reservoir component, that is, oil, connate water,
voir fluids allows achieving a comprehensive energy-bal-
and formation rock, and combine all parameters into
ance equation of the system. The general form of thermal
Eq. (14) to obtain
energy balance, in terms of the equation of change for
internal energy, was presented in Eq. (1). This equation can   oT
£so qo cpo þ £sw qw cpw þ ð1  £Þqf cpf þ qo ur cpo
also be written in terms of enthalpy, temperature, and ot
pressure. In other words, Eq. (1) can be rewritten in the oT op op
þ qo ur ro þ ½£so qo ro þ £sw qw rw  1
following form: or   or ot
    1o oT
o op * * * ¼ kr þ Q_ ð17Þ
qH  ¼  r  qHu þ r  pu  r  q r or or
ot ot    
 s : ru  p r  u þ Q_
* * *
ð11Þ Equation (17) or Eq. (2) in the text is fundamentally the
same as the thermal energy-balance equation presented by
For the 1D radial system, the double-dot product can be App (2009, 2010). This equation is also the basis for our
represented by Newton’s law of viscosity, analytical model to evaluate the flowing-fluid temperature
 
ou 2
* *
 s : ru ¼ 2l or , and heat conduction can be
r distribution in the reservoir.
*
  Let us list the underlying assumptions of this model.
represented by r  q ¼ 1r oro kr oT or ; therefore, Eq. (11) This study presupposes that the reservoir is homogeneous,
can be rewritten as and the rock and fluid properties are time invariant. Other
   2 general assumptions include the following:
o 1o 1o oT our Dp
qH þ ½rqHur  ¼ kr þ 2l þ
ot r or r or or or Dt 1. The only flowing fluid in the reservoir is oil.
þ Q_ 2. The reservoir is producing at a constant rate.
ð12Þ 3. The original temperature of overburden and under-
burden formations is the same as the reservoir
Al-Hadhrami et al. (2003) examined the viscous temperature at initial conditions. The elevation
dissipation term in the energy-balance equation in differences from reservoir depth are negligible.
Cartesian coordinates. Following their work, we 4. Overburden and under-burden formations are infinite
approximated the viscous term for the 1D radial flow in sources/sinks. Overburden and under-burden
porous media with ur op
or . Equation (12) then becomes

123
808 J Petrol Explor Prod Technol (2018) 8:799–811

formations remain at their original temperatures even formation undisturbed temperature in a complex manner.
after heat transfer to/from the reservoir occurs. We approximate this term by following App’s (2010)
5. Radial heat conduction is negligible during constant approach using Newton’s law of cooling, giving
rate production. Q_ ¼ 2hc ½T  Ts =h, where hc is the heat transfer
6. The pressure transient term, qp/qt, is assumed to be coefficient. The complete energy-balance equation of the
negligible. Therefore, for a given flow rate, pressure system then becomes
varies in the radial direction, but not with time.   oT
7. Fluid temperature and pressure remain constant at the £so qo cpo þ £sw qw cpw þ ð1  £Þqf cpf
ot
reservoir boundary. ð22Þ
8. Porosity and permeability remains unchanged. qqo cpo oT lq2 qo ro 2hc ½T  Ts 
  ¼ 
9. Variation in fluid properties of density and viscosity 2prh or ð2prhÞ2 k h
is negligible.
If we use lumped parameters A, B, and C, Eq. (21) can
10. The fluid’s local velocity (superficial velocity) can be
be simplified to the following expression:
estimated from Darcy’s equation:
oT oT 4pr 2 hc 4pr 2 hc
kA op 2prhk op Ar 2  Br C ¼ Tþ Ts ð23Þ
q¼ ¼ ð18Þ ot or q q
l or l or
Equation (14) is a first-order, partial-differential
q q k op equation where fluid temperature T is a function of radial
ur ¼ ¼ ¼ ð19Þ
A 2prh l or distance r from wellbore into the reservoir and producing
Note that these assumptions are necessary to obtain a time t. Initially, fluid temperature in the reservoir is
useful analytical solution to the flow problem at hand; the constant at Ti. We apply the method of characteristics with
model validation section explores the effects of some the initial condition, T ðr; t ¼ 0Þ ¼ Ti to arrive at the
simplifying assumptions on solution quality. following solution to Eq. (23):
 
C HðAr2 þ2BtÞ H ðAr 2 þ 2BtÞ
T ðr; tÞ ¼ Ti þ e 2B Ei 
2B  2B
Appendix B: Analytical solution for reservoir
2  2
C HAr HAr
flowing-fluid temperature estimation  e 2B Ei  ; ð24Þ
2B 2B

A comprehensive energy-balance equation for the system where



with a consideration of heat transfer between the system   2ph
and surroundings is given by Eq. (18). Based on our gen- A ¼ £so qo cpo þ £sw qw cpw þ ð1  £Þqf cpf
q
eral assumptions, radial heat conduction during constant
ð25Þ
rate production is negligible. Additionally, qp/qt term is
assumed to be minimal and can be omitted. Therefore, B ¼ qo cpo ð26Þ
Eq. (17) becomes qq ro l
C¼ o ð27Þ
  oT 2phk
£so qo cpo þ £sw qw cpw þ ð1  £Þqf cpf
ot ð20Þ 4hc p
oT op D¼ ð28Þ
þ qo ur cpo _
þ qo ur ro ¼ Q q
or or
4hc p
E¼ Ti ð29Þ
We rewrite velocity ur in terms of flow rate as q/(2prh) q
and rewrite the qp/qr in terms of flow rate as -(lq)/(2prhk).
D
Also, we replace q with –q because flow occurs in the H¼ ð30Þ
A
negative r-direction during production. Therefore, Eq. (20)
becomes Note that parameters A through H are constant for a
  oT particular reservoir and are reported in the main text as
£so qo cpo þ £sw qw cpw þ ð1  £Þqf cpf Eqs. (5)–(10); Chevarunotai (2014) presents further details
ot
ð21Þ of this derivation.
qqo cpo oT lq2 qo ro _
  ¼ Q
2prh or ð2prhÞ2 k

The net input rate of energy between reservoir and


under- and overburden formations Q_ is related to the

123
J Petrol Explor Prod Technol (2018) 8:799–811 809

10
Table 1 Reservoir static parameters (after App 2010) 318 °F

Parameter, units Value 300 °F

k, md 20

Oil Viscosity, cp
/, % 25
h, ft 100
pi, psia 21,000
pb, psia 7000
ct, psi-1 3 9 10-6
Ti, °F 302 1
5,000 7,000 9,000 11,000 13,000 15,000 17,000 19,000 21,000 23,000
rw, ft 0.41 Pressure, psia
re, ft 4000
Swi, % 15 Fig. 8 Oil viscosity as a function of pressure and temperature (after
2 App 2010)
hc, BTU/hr ft °F 0.92

Appendix C: Input data for model validation Yoshioka (2013) has shown that the effect of formation
thermal conductivity can be represented by the Peclet
We used a real field example, reported previously in App’s number. Relating fluid velocity u to production rate q, App
study (2010). App had generated his solutions with a newly and Yoshikawa expressed Pe as follows:
developed numerical model. Table 1 shows static (rock) ur qq cpo
properties of the reservoir, which is considered as a base Pe ¼ ¼ o ð31Þ
a 2phk
case in this study. We assumed that the reservoir is
homogeneous and that the model parameters remain con- Their work showed that when 0.1 [ Pe [ 3, the effect
stant throughout the production period. Table 2 presents of Pe, and, therefore, formation thermal conductivity k on
the reservoir fluid properties. Oil is the flowing phase, and fluid temperature change due to expansion, is negligible.
the formation water is considered immobile and that no The calculations for Pe for our lowest production rate of
production of free gas occurs in the reservoir, given the 970 STB/D are
!
low-saturation pressure. STB ft3
5:615 STB
Another critical fluid property in flowing-fluid temper- q ¼ 970  ¼ 227ft3 =hr ð32Þ
D 24 hr
D
ature and well productivity calculation is oil viscosity. Data
3
from laboratory measurements for this particular reservoir qqcp 227 fthr  51:19 lbm Btu
 0:53 lbm 
ft3   F ¼ 5:6
were also presented in App’s paper. Figure 8 presents the Pe ¼ ¼ ð33Þ
2phk BTU
2p  100 ft  1:73 hr  F ft
oil viscosity as a function of pressure and temperature,
which is used for model validation.
Similarly, for our highest production rate of 6200 STB/
An important assumption in our model is that conduc-
D, Pe = 36.3. Therefore, for the cases considered in this
tive heat flow for our problem is negligible. App and
study, omitting thermal conductivity of the formation does
not introduce any significant error. We note that for most
Table 2 Reservoir fluid parameters (after App 2010) deepwater assets, economic production rates are expected
to be high enough to result in correspondingly high Peclet
Parameter, units Value
numbers. As a consequence, the underlying assumptions
Bo, bbl/STB 1.05 made while deriving the analytical formulation of this
3
qo, lbm/ft 51.19 coupled fluid and heat flow problem appear reasonable.
cpo, BTU/lbm/°F 0.53
ro, °F/psi - 0.0055
qw, lbm/ft3 63.68 Appendix D: Gas properties
cpw, BTU/lbm/°F 1.0
rw, °F/psi - 0.0024 To make our model suitable for gas temperature calcula-
qf, lbm/ft 3
165.43 tion, we allowed several properties to vary. The models or
cpf, BTU/lbm/°F 0.20 correlations used for these computations are presented
k, BTU/hr ft °F 1.73 below.
l, cp 2.0

123
810 J Petrol Explor Prod Technol (2018) 8:799–811

Gas viscosity Table 3 Reservoir and fluid properties (App 2009)


Parameter Value Parameter Value
Gas viscosity correlations have been presented by a num-
ber of authors. For our calculations, we used the one by Lee k, md 3.5 Ti, °F 376
et al. (1966): Porosity, % 15 rw, ft 0.41

Thickness, ft 20 re, ft 4000
lg ¼ A 104 EXPðBqcg Þ ð34Þ
pi, psia 3880 Sw, % 15
where pd, psia 650 Bg@pi, rb/mcf 1.063
cr, psi-1 3E-6 cw, psi-1 3E-6
ð9:379 þ 0:01607 Ma ÞT 1:5
A¼ ð35Þ
209:2 þ 19:26 Ma þ T
986:4 Table 4 Component thermal and physical properties
B ¼ 3:448 þ þ 0:01009 Ma ð36Þ
T
Properties Gas Water Rock
C ¼ 2:447  0:2224B ð37Þ
3
q, lbm/ft 11.58 62.43 165.0
Gas density cp, Btu/lbm °F 0.52 1.00 0.23
b, 1/°F 4E - 4 5E - 4 5E - 4
The gas density is calculated from real gas law, as follows: lJT, °F/psi - 0.0024
ZWRT kt, Btu/hr ft °F 1.6
pV ¼ ZnRT ¼ ð38Þ
Ma
The Newton–Raphson approach is used to solve for Z
pMa 0:00149406 Ma
oZ
q¼ ¼ ð39Þ and oT .
ZRT RT p

where q is gas density in gm/cc, Ma is apparent molecular


weight, T is the temperature in °R, Z is the gas law devi-
Appendix E: Properties for low-pressure gas DST
ation factor, and p is the pressure in psia.
The following properties are taken from App (2009) for
Joule–Thomson coefficient
temperature analysis of a low-pressure gas production test
(Tables 3, 4).
An expression for calculating Joule–Thomson coefficient
for real gases has been developed by Hasan et al. (2010):
   
VT oZ VT oZ
cp CJ ¼ V  V  ¼ ð40Þ
Z oT p Z oT p References

Equation (40) requires accurate estimates of Z and Al-Hadhrami AK, Elliott L, Ingham DB (2003) A new model for

oZ viscous dissipation in porous media across a range of perme-


oT p . We use the following expression for these by ability values. Transp Porous Media 53(1):117–122. doi:10.
Dranchuk et al. (1973): 1023/A:1023557332542
App JF (2009) Field cases: nonisothermal behavior due to Joule–
Z ¼ 0:27pr =Tr q ð41Þ Thomson and transient fluid expansion/compression effects. In:
Paper SPE-124338-MS presented at the 2009 SPE annual
In Eq. (41), q is a polynomial function of pr (= p/pc) and
technical conference and exhibition, New Orleans, Louisiana,
Tr (= T/Tc), which is given by 4–7 Oct 2009. SPE-124338-MS. doi:10.2118/124338-MS

2 App JF (2010) Nonisothermal and productivity behavior of high-
f ðqÞ ¼ aq6 þ bq3 þ cq2 þ dq þ eq3 1 þ f q2 ef q  g pressure reservoirs. SPE J 15(1):50–63. doi:10.2118/114705-PA
ð42Þ App JF, Yoshioka K (2013) Impact of reservoir permeability on
flowing sandface temperatures: dimensionless analysis. SPE J
The constants are as follows: 18(4):685–694. doi:10.2118/146951-PA
a = 0.06423, b = 0.5353Tr - 0.6123, c = 0.3151Tr - Bird RB, Stewart WE, Lightfoot EN (2006) Transport phenomena,
2nd edn. Wiley, New York, p 336. ISBN 978-0470115398
1.0467 - 0.5783/T2r , d = Tr, e = 0.6816/T2r , f = 0.6845 Chevarunotai N (2014) Analytical model for flowing-fluid tempera-
g = 0.27pr ture distribution in single-phase oil reservoir accounting for
Joule–Thomson effect. MS Thesis, Petroleum Engineering

123
J Petrol Explor Prod Technol (2018) 8:799–811 811

Department, Texas A&M University, College Station, Texas, Oudeman P, Kerem M (2006) Transient behavior of annular pressure
USA build-up in HP/HT wells. SPE Drill Complet 21(4):234–241.
Dawkrajai P, Lake LW, Yoshioka K et al (2006) Detection of water or doi:10.2118/88735-PA
gas entries in horizontal wells from temperature profiles. In: Ramazanov AS, Nagimov VM, Akhmetov RK (2013) Analytical
Paper SPE 100050 presented at the 2006 SPE/DOE Symposium model of temperature prediction for a given production history.
on Improved oil Recovery, Tulsa, Oklahoma, 22–26 April 2006. Oil Gas Bus J (1):537–546. http://ogbus.ru/eng/authors/
doi:10.2118/100050-MS Ramazanov/Ramazanov_4e.pdf
Dietz DN (1965) Determination of average reservoir pressure from Satman A, Brigham WE, Zolotukhin AB (1979) A new approach for
build-up surveys. J Pet Technol 17(8):955–959. doi:10.2118/ predicting the thermal behavior in porous media during fluid
1156-PA injection. Geotherm Resour Counc Trans 3:621–624
Dranchuk PM, Purvis RA, Robinson DB (1973) Computer calculation Shook GM (2001) Predicting thermal breakthrough in heterogeneous
of natural gas compressibility factors using the Standing and media from tracer tests. Geothermics 30(6):573–589. doi:10.
Katz correlation. In: Annual technical meeting, Petroleum 1016/S0375-6505(01)00015-3
Society of Canada, May 8–12, Edmonton, Canada. doi:10. Spillette AG (1965) Heat transfer during hot fluid injection into an oil
2118/73-112 reservoir. J Can Pet Technol 4(4):213–218. doi:10.2118/65-04-
Duru OO, Horne RN (2010) Modeling reservoir temperature 06
transients and reservoir-parameter estimation constrained to the Steffensen RJ, Smith RC (1973) The importance of Joule–Thomson
model. SPE Res Eval Eng 13(6):873–883. doi:10.2118/115791- heating (or cooling) in temperature log interpretation. In: Paper
PA SPE-4636-MS presented at the SPE-AIME 48th annual fall
Hasan R, Izgec B, Kabir CS (2010) Sustaining production by meeting, Las Vegas, Nevada, 1–3 Oct 1973. doi:10.2118/4636-
managing annular-pressure buildup. SPE Prod Oper MS
25(2):195–203. doi:10.2118/120778-PA Tan H, Cheng X, Guo H (2012) Closed solutions for transient heat
Izgec B, Kabir CS, Zhu D et al (2007) Transient fluid and heat flow transport in geological media: new development, comparisons,
modeling in coupled wellbore/reservoir systems. SPE Res Eval and validations. Transp Porous Media 93(3):737–752. doi:10.
Eng 10(3):294–301. SPE-102070-PA. doi:10.2118/102070-PA 1007/s11242-012-9980-5
Kabir CS, Izgec B, Hasan AR et al (2012) Computing flow profiles Yoshioka K, Zhu D, Hill AD et al (2005) A comprehensive model of
and total flow rate with temperature surveys in gas wells. J Nat temperature behavior in a horizontal well. In: Paper SPE-95656-
Gas Sci Eng 4:1–7. doi:10.1016/j.jngse.2011.10.004 MS presented at the 2005 SPE annual technical conference and
Lauwerier HA (1955) The transport of heat in an oil layer caused by exhibition, Dallas, Texas, 9–12 Oct 2005. doi:10.2118/95656-
injection of hot fluid. Appl Sci Res 5(2–3):145–150 MS
Lee AL, Gonzalez MH, Eakin BM (1966) The viscosity of natural Yoshioka K, Zhu D, Hill AD et al (2006) Detection of water or gas
gases. Trans AIME 237:997–1000 entries in horizontal wells from temperature profiles. In: Paper
Onur M, Cinar M (2016) Temperature transient analysis of slightly SPE-100209-MS presented at the SPE Europec/EAGE annual
compressible, single-phase reservoirs. In: Paper SPE-180074- conference and exhibition, Vienna, Austria, 12–15 June 2006.
MS presented at the 78th EAGE conference and exhibition, doi:10.2118/100209-MS
Vienna, Austria, 30 May–2 June 2016. doi:10.2118/180074-MS
Onur M, Palabiyik Y (2015) Nonlinear parameter estimation based on
history matching of temperature measurements for single-phase Publisher’s Note
liquid–water geothermal reservoirs. In: Proceedings, world
geothermal congress, Melbourne, Australia, 19–25 April 2015. Springer Nature remains neutral with regard to jurisdictional claims in
https://pangea.stanford.edu/ERE/db/WGC/papers/WGC/2015/ published maps and institutional affiliations.
22009.pdf

123

Anda mungkin juga menyukai