Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

PROCEEDINGS, Fortieth Workshop on Geothermal Reservoir Engineering

Stanford University, Stanford, California, January 26-28, 2015


SGP-TR-204

Analysis of Wellhead Production Temperature Derivatives


Kaan Kutun, Omer Inanc Tureyen, Abdurrahman Satman
Istanbul Technical University Maslak 34469 Istanbul Turkey
kutunk@itu.edu.tr, inanct@itu.edu.tr, mdsatman@itu.edu.tr

Keywords: Geothermal, wellhead temperatures, derivatives.

ABSTRACT
It is commonplace to take pressure and temperature measurements along the wells in geothermal reservoirs. Such profiles give good
insight into the characterization of the geothermal reservoir. Many useful information such as upper and lower boundaries of reservoirs,
deliverability of wells, the nature of the phase changes within the wells can be obtained as a result of these profiles. However in most
cases taking these profiles along the wells can be challenging. Furthermore these types of profiles cannot be considered at all times since
the wells would have to be shut in and there is always the economical factor.

In this study we look at in more detail the behavior of the wellhead temperature, the only place where temperature can be measured at
all times and is not an economical burden. We specifically study the behavior of the temperature derivative obtained at the wellhead.
Just as in a well testing diagnostic plot the temperature derivative at the wellhead displays clearly various flow regimes that take place in
the well. At early times when wellbore convection dominates, a log-log plot of dT/dlnt versus t gives a unit slope straight line similar to
early-time pressure transient behavior dominated by wellbore storage. The intermediate-time behavior is observed when heat transfer
between wellbore and the surrounding formations is felt and is recognized by -1/2 slope straight line on a log-log plot of dT/dlnt versus
t. At late times the wellbore heat transmission solution converges to the cylindrical-source solution transferring heat at constant
temperature.

1. INTRODUCTION
Generally speaking, methods and models describing the wellbore heat transmission is based on a wellbore heat balance with some
assumptions. The most critical assumptions refer to the condition between the wellbore and the surrounding formations. Either a
constant heat flux or a constant temperature assumption is used for this purpose. In the case of the general wellbore heat problem,
neither heat flux nor the temperature at the wellbore remains constant except in special cases. However, the solutions for the cases of
sources loosing heat at constant flux and constant temperature eventually converge at long times.

Most of the literature on wellbore heat transmission is based on the classical work by Ramey (1962). He derived the temperature
distribution in a well used for injecting hot fluid. Ramey (1964) expanded on this to estimate the rate of heat loss from the well to the
formation. Horne (1979) reexamined the problem to determine the wellbore heat loss in production and injection wells. Hagoort (2004)
assessed Rameys classical method for the calculation of temperatures in injection and production wells and showed that Rameys
method is an excellent approximation.

Carslaw and Jaeger (1959) present graphical and analytical solutions for the cases of internal cylindrical sources losing heat at constant
flux, constant temperature and the radiation boundary conditions. Solutions converge at long times. This is a sufficiently long time at
which temperature is controlled by formation conditions.

For small values of time heat flow in the wellbore is controlled by convection, rather than conduction. Ramey recommends using the
constant-temperature cylindrical-source solution if thermal resistance in the wellbore is negligible that is the case when the fluid flow
occurs thru casing only.

Our study here considers only single-phase fluids flowing in the well. The single-phase flow analysis is based on the determination of
the fluid temperature as a function depth and time. For simplicity, most of our work here assumes a geothermal well with single-phase
liquid flowing in a casing without tubing, however it is not difficult to derive the equations and give the expressions for fluid flow
within tubing case.

Rameys solutions for temperature are obtained in terms of depth. However the effects of varying formation temperature as a function of
depth in terms of geothermal temperature gradient as well as heat transfer to the surrounding formations by transient conduction are
considered in those solutions.

The equation for the evaluation of the temperature in a producing geothermal well is (Ramey, 1962):

= ( ) + (1 ) (1)

1
Kutun et al.

where y is the distance upwards from the bottom of the well, Tbh is the downhole reservoir temperature, (Tbh-y) is the temperature of
the earth (Te) assuming linear geothermal gradient, and is the geothermal gradient (the increase of formation temperature with increase
in depth).

A is a group of variables defined as


()
= (2)
2

where w is the mass flow rate, c is the thermal heat capacity (specific heat) of the fluid (assumed constant), k is the thermal conductivity
of the formation (earth), and f(t) is a dimensionless time function representing the transient heat transfer to the formation.

The function f(t) may be found from Ramey (1962 and 1964), or alternatively may be approximated from the line source solution for
flowing times greater than 30 days or so by the equation (Rameys large time solution):

() ( ) 0.29 (3)
2

where is the thermal diffusivity of the formation and rw is the radius of the casing.

Heat transfer from the casing to the formation in terms of heat flow rate per unit of length is given by:
2
= ( ) (4)
()

where T is the temperature of the fluid in the casing and Te is the temperature of the formation. Thus the total heat-loss rate from a well
of total depth L can be given as (Horne, 1979)

= [ + ( 1)] (5)

Application of the Ramey equation for the wellbore fluid temperature requires as input f(t) function at a certain time. Various studies in
the literature propose expressions to estimate f(t). These expressions are given in the following discussion.

1.1 f(t) Correlations For A Cylindrical Source With Constant Temperature and With Constant Heat Flux
1.1.1 A circular cylinder with constant heat flux at the surface case
It is well-known that in this case the formulation has a solution in complex integral form (Van Everdingen and Hurst (1949); Carslaw
and Jaeger (1959)).

For large values of the time: For large values of the time, Carslaw and Jaeger (1959)s solution yields the following expression for a
circular cylinder of radius rw, with constant heat flux at the surface (r=rw):

1 4
= ( ) (6)
2 2

where


= 2 (7)

and C=e=e0.57722=1.781.

Rameys heat flux solution is given by Eq. 4. Comparison of Eqs. 6 and 4 gives:
1 4
() = ( ) = (2.246 ) = ln(1.4986 ) (8)
2

For small values of the time: Kutasov (2003) indicates that for an infinite cylindrical source of heat the following expression
represents f(t) at small values of tD:


() = 2 (9)

1.1.2 A circular cylinder with constant bore-face temperature case


Carslaw and Jaeger (1959) gave a solution for this case. However the solution is in a complex integral form and thus it is not repeated
here. Later on Ramey (1962, 1964, and 1981) presented f(t) results in graphical and tabulated forms. Kutasov (1987) and Kutasov and

2
Kutun et al.

Eppelbaum (2005) used a semi-theoretical approach to obtain the following expression to represent f(t) for the case of a flow in a
wellbore with constant wellbore temperature at the bore-face:
1
() = [1 + (1.571 ) ] (10)
4.959+

For large values of tD, Eq. 10 reduces to:

() = ln(1.77 + 1) (11)

In this study we propose the following simplified expression for f(t):

() = (1 + 1.7 ) (12)

This equation is based on a best curve fit of Rameys f(t) data (Ramey; 1962, 1964, and 1981) in the pertaining time interval and is
accurate within 1% as will be discussed later. Furthermore for values of tD>20 the following expression within about 5% accuracy can
be used:

() = (1.7 ) (13)

Satman et al. (1979, 1984) discussed the time-dependent overall heat-transfer coefficient concept to heat-transfer problems and
introduced the following expression to represent f(t):

() = (14)

Notice that Eq. 14 is identical in form to Eq. 11 given by Kutasov and Eppelbaum (2005) after modification for small values of tD using
the simplification of ln(1+x)=x.

Figures 1 and 2 compare the various f(t) correlations discussed above. The constant heat flux line source model yields an f(t) being equal
to (1/2)Ei(1/4tD) where Ei(x) is the exponential integral of x. The numerical data representing the constant temperature at cylindrical
source model in Figures 1 and 2 are taken from Ramey (1981).

Figure 1: Comparison of f(t) correlations on a semilogarithmic graph.

A detailed analysis of the correlations given in Figures 1 and 2 indicates that Eq. 12 for all times, Eq. 9 for early times, and Eq. 13 for
late times proposed in this study are quite adequate for most engineering purposes. Equations 10 and 12 match the dimensionless time
function curve for cylindrical source with a constant temperature at wellbore presented by Ramey (1962, 1964, and 1981), Earlougher
(1977) and Hagoort (2004). The match is acceptable for practical engineering calculations and Eqs. 12 and 13 have much easier forms
to be used.
3
Kutun et al.

Figure 2: Comparison of early- and late-time f(t) correlations on a semilogarithmic graph.

One particular observation about the heat transfer models compared in Fig.1 is that the line source with a constant heat flux yields
lower f(t)s than the cylindrical source with a constant temperature, although they eventually converge at long times at about tD=105
(Hagoort, 2004). The difference in f(t)s is considerable, about 50%, at tD=1 and becomes lower as tD grows, about 5% at tD=1000.

1.2 Temperature Solution At Early-Time Wellbore Convection Dominated Period


The parameter A in Eq. 1 expresses the ratio of the heat transported up the wellbore by convection to the heat loss by conduction in the
formation. For large values of A (high production rates), the heat transport is dominated by convection and little heat will be lost to the
formation.

Rameys solution (Eq. 1) containing the parameter A for the temperature distribution is based on the following two simplifying
assumptions in the wellbore heat balance: (i) the initial accumulation term up to the filling of the welllbore is negligible, and (ii) the heat
loss term can be replaced by the product of the f(t) function and temperature. The first assumption definetely does not hold during the
early stages when the hot fluid entering the wellbore at the bottom is displacing the initial wellbore fluid. In this period, the temperature
profile in the wellbore is coupled to the displacement front, the movement of which is controlled by the accumulation up to the filling of
the wellbore.

Using rigorous solution of the wellbore and formation heat-balance equations Hagoort (2004) proposed the following early-time
solution for the temperature ahead of the fluid front ( (y/L)>(ut/L));

= + (15)

where u is the linear velocity in the wellbore, and L is the wellbore length. Early means a time long before the fluid front reaches the
wellbore. Equation 15 reflects the displacement of the initial fluid with the linear temperature profile from the wellbore, the profile
ahead of the front moves up linearly proportional to time.

Hagoort (2004) also presented the following early-time solution (ahead of the displacement front, (y/L)>(ut/L) ) based on heat balance
equations assuming constant heat-loss function:

= + {1 (/)} (16)

Equation 16 represent the wellbore temperature distribution before the arrival of the front at the wellhead for t<L/u.

Comparing the rigorous solutions obtained from the heat balance model, Hagoort (2004) concluded that Rameys method is an
excellent approximation, except for an early transient period in which the calculated temperatures are significantly overestimated.

1.3 General Characteristics of Production Temperature Profiles


Figure 3 presents a comparison of temperatures obtained in a water-flowing well with temperatures computed for production a
conditions. The thermal properties used in modeling the temperature profiles and related discussions are given in Table 1.

4
Kutun et al.

Table 1-Input parameters for the cases discussed

Earth temperature at surface, oC 20


Rock density, kg/m3 1952
Formation thermal conductivity, J/m.s.oC 2.92
Formation specific heat capacity, J/kg.oC 1000
Formation thermal diffusivity, m2/D 0.129

The water-production rate is 20 kg/s, the bottomhole reservoir temperature is 245 oC, the surface (earth) temperature is 20 oC, and the
wellbore radius is 0.15 m. Figure 3 illustrates the general features of the wellbore temperature distribution. The static temperature
profile (the geothermal gradient) represents the formation temperature as a function of depth for undisturbed wellbore conditions.

The dynamic temperature profiles for the production case are plotted at four successive times; 0.1, 1, 10, and 30 days. The potential
importance of time on wellbore heat transmission is shown by Fig. 3.

Figure 3: Temperature profiles at successive times for production case.

As seen in Figure 3, the dynamic production profile for the smallest time of 0.1 day is concave, reflecting the greater effect of heat
losses from wellbore to surrounding formations at shallower depths. Because of continuous heating of the formation by the hot water in
wellbore, heat losses become progressively smaller with time. Therefore, the temperature profile moves to the right, closer to the inlet
bottomhole temperature.

Figure 3 shows how the temperature profile inside the well evolves with time. It is important to note that, most of the change occurs
during the first 10 days. Then the changes in profile with time become very small. Two main mechanisms of heat transfer play a role in
the changing well temperature profile. The first is the convective heat transfer from the bottom of the well to the top via production of
the fluid. The second is the conductive heat loss to the surroundings of the well. In the final profile at 30 days, the wellhead temperature
gets closer to the initial reservoir temperature. This difference between the wellhead temperature and the bottomhole temperature is
because of the heat losses to the surroundings of the well.

The effect of time on temperature profiles are clearly demonstrated in Fig. 2. As production time increases, the temperature at a given
depth increases. The change in temperature at a given depth with time is very rapid at first, then decreases markedly. Heat transfer rate
at early times is controlled by convection in wellbore but at large times it is controlled by the rate at which heat is conducted within the
surrounding media.

2. WELLHEAD TEMPERATURE DERIVATIVE BEHAVIOR


The general theory on wellbore heat transmission is already discussed in previous sections. Figure 4 schematically illustrates the
general characteristics of the wellhead temperature (Ttop) as a function of time. The temperature behavior consists of four time periods;
an early-time period, a transition period, an intermediate-time period, and a late-time period. In Fig. 4, T corresponds to the difference

5
Kutun et al.

between the wellhead temperature Ttop and the surface earth temperature Tsurf. It is determined from Eq. 1 which can be written in terms
of the Ramey Number, NRa=2kL/(wc):

( )
= () [1 ( )] (17)
()

where Tbot is the bottom-hole temperature. In the early-time period the cylindrical source constant temperature solution (Ramey; 1962,
1964, 1981) represented by the dashed-curve overestimates the wellhead temperature. Equation 15 is used to determine the wellhead
temperature at early-time period and Hagoort (2004) indicates that the duration of this period depends on /( 2 ); the larger the
/( 2 ), the longer the early-time period. A transition period exists after the early-time period and the wellhead temperature solution
converges with the cylindrical source constant temperature solution about one and half log cycle after the wellbore convection
dominated early-time period which is diagnosed by a unit slope straight line relationship on a log T-log t graph to be discussed in the
following section. The wellhead temperature behavior in the intermediate-time period is represented by Eq. 1 with a dimensionless time
function f(t) described by an expression proportional to such as Eqs. 9 and 14. The duration of this period depends on /( 2 );
the larger the /( 2 ), the shorter the intermediate-time period.

Throughout the time period in which the cylindrical source solution is valid Eq. 1 describes the wellhead temperature (or temperature at
any y) employing Eq. 12 for f(t). In the late-time period, Eq. 13 is used an approximation and simplification of f(t) for tD>20.

Cylindrical Source Constant Temperature


Ramey (1962, 1964, 1981)

Early-Time
Wellbore
Convection
Dominated Late-Time Period
Period

tD>20.0

Intermediate-
Unit Slope
Time Period,
Transition Period
~1.5 log cycle

t<L/u
t<L/(ur2)

Figure 4: General characteristics of the wellhead temperature.

In the following sections the effect of time on the temperature behavior in terms of temperature derivatives is assessed.

2.1 Early-Time Temperature Derivative Behavior


For the temperatures measured at the wellhead (y=L), Eq. 16 which represents the wellbore temperature profile before the arrival of the
displacement front at the wellhead is employed. Eq. 16 can be modified to:

= = {1 (/)} (18)

Using the relationship exp(-x)1-x since x is too small, Eq. 18 reduces to

= (19)

Eq. 19 is identical to Eq. 15 given by Hagoort (2004).

Now we can take derivatives of T with respect to time t in terms of dT/dt and dT/dlnt;

6
Kutun et al.

dT/dt=u (20)

and

dT/dlnt=ut (21)

Equation 20 indicates that the time derivative of wellhead temperature change stays at a constant value, u, and moreover Eq. 21 shows
that a log-log plot of dT/dlnt (or tdT/dt) versus t gives a unit slope straight line.

The early-time temperature behavior is similar to the wellbore storage phenomenon occuring in oil and gas production wells and
affecting short-time transient pressure behavior (Earlougher, 1977). The temperature behavior in geothermal wells and the pressure
behavior in oil and gas wells are both due to the liquid stored in the wellbore when the liquid level rises.

Equations 19 and 21 both have a characteristic that is diagnostic of wellbore convection dominated period: the slope of the T and
dT/dlnt versus t graphs on log-log paper is 1.0 during wellbore convection domination, respectively. Temperature-time data falling on
the unit slope of the log-log plot reveal nothing about the surrounding formation properties since all production is from the wellbore
during that time.

Note that the location of the log-log unit slope can be used to estimate the value of u from Eq. 19. T and t are values read from a point
on the log-log unit slope straight line. u is calculated from Eq. 19.

2.2 Intermediate-Time Temperature Derivative Behavior


Using exp(-x)1-x+x2, the Rameys wellbore heat transmission equation, Eq. 1, for the wellhead temperature reduces to the following
intermediate-time temperature solution valid after the displacement front reaches the wellhead:
2
2
=
(22)
2()

Our study indicates that the early time f(t) correlation, Eq. 9, given by Kutasov (2003) matches the Rameys f(t) values within the
reasonable accuracy for the time range after the convection dominated period. Thus assuming the f(t) correlation, Eq. 9, is valid the
derivatives of T in terms of dT/dt and dT/dlnt are obtained from Eq. 22 as following:
2
2 3
=
2 (23)

4.5 2

2
2 1
=
2 (24)

4.5 2

Notice that the plot of log(dT/dt) vs. log(t) gives a straight line with a slope -3/2 whereas the plot of log(T/dlnt) vs. log(t) gives a
straight line with a slope -1/2.

One further observation regarding the wellhead derivative solutions, equating the early-time and the intermediate-time wellhead
temperature derivative relationships, Eqs. 20 and 23, can give the intersection time of the early-time and intermediate-time straight lines
(tei) as
2/3
1.42 2 3 2/3
= [ ] = 2.68 [ ] (25)
2
2

The intermediate-time behavior described by Eqs. 22-25 is observed when the effect of the heat transfer between the wellbore and the
surrounding formation becomes significant after the early-time convection dominated period. It could be of a short duration and even
may not exist depending on the scales of the heat transfer rate (q) and the wellbore convective flow. It can be observed if the the
wellbore convective flow is relativeley small and the early-time period is long and the conditions for f(t) valid for small values of time
(Eq. 9) is dominant.

The validity of the temperature derivative solutions given above are checked using the wellbore temperature model described by Kutun
et al.(2014, 2015). The thermal properties used in modeling are given in Table 1. The well dimensions, the flow rate, and other relevant
data are given in Table 2. In all the synthetic applications, presented in this section, the parameters given in Table 1 and 2 are used
unless otherwise stated.

7
Kutun et al.

Table 2- Input parameters for the cases discussed

Well depth, m 1050.0


Well radius, m 0.1
Bottom hole temperature, oC 144.09
Water production rate, kg/s 4.39
Geothermal temperature gradient, oC/m 0.118
Production time, days 50.0

Figures 5-8 give T, dT/dt, and dT/dlnt versus time (t) plots obtained from the wellbore simulator by Kutun et al. (2014). Figure 5
shows the temperature derivative results for a 50 day production period whereas Fig. 6 illustrates the intermediate-time temperature
derivative with a -3/2 slope. Similarly Fig. 7 gives dT/dlnt versus time on a log-log plot whereas Fig. 8 exhibits the unit slope and -1/2
slope periods on the same plot. Figures 7 and 8 show T versus t and dT/dlnt versus t on the same plot for comparison purposes.

Figure 5: dT/dt versus time plot.

-3/2 Slope

Figure 6: dT/dt versus time plot and -3/2 slope.

8
Kutun et al.

Figure 7: T versus time and dT/dlnt versus time plots.

Unit Slope

Half (-1/2) Slope

Figure 8: T versus time and dT/dlnt versus time plots and unit slope and -1/2 slope characteristics.

For this particular data the linear velocity in the wellbore u is 0.14 m/s and the fluid front reaches the wellbore at 0.087 day. As seen in
Figs. 5-8 the convection dominated period with a unit slope ends at about 0.003 day which indicates the validity of the condition t<L/u.
Thus the earlier temperature data represents the convection dominated period.

Figure 7 shows T versus t on log-log plot. As indicated earlier it is possible to estimate u from the unit-slope straight line. At
t=0.0001 day, T=0.14 oC on the log-log unit-slope straight line so that T/t is 1400 oC/day that is being equal to u as Eq. 19 indicates.
The input value for u is 1424 oC/day which matches the calculated value perfectly well.

9
Kutun et al.

The intermediate-time period for which the intermediate-time behavior and the cylindirical source solution as well are expected to begin
at approximately one and half log cyle after the end of the early-time period (0.003 day) that is 0.06 day. That is the case as observed in
Fig. 8.

Figure 8 shows that the early-time unit-slope straight line and the intermediate time -1/2 slope straight line intersect at about 0.03 day.
Using the input data in Eq. 25, the intersection time tei is determined to be 0.032 day that validates the accuracy of Eq. 25.

2.3. Late-Time Temperature and Heat-Transfer Behavior

Theoretically the heat transmission characteristics in a borehole heat exchanger and in a geothermal well are similar. The method of
estimating thermal conductivity from borehole heat exchanger temperature measurements is applied provided that the heat transfer rate
per unit length (q) is kept at a constant value.
For a geothermal well with flow in casing so that the thermal resistance term (Rb) is neglected, the total heat gained per length is
expressed by Eq. 4. Writing Eq. 4 in terms of the wellhead temperature and using Eq. 13 for f(t) yields:

(1.7 2 )
1
= () + + (26)
4 2

Assuming a constant heat transfer rate per length (q), then the plot of wellhead temperatures measured versus the natural logarithm of
time is expected to give a semilog straight line with a slope being equal to m and the thermal conductivity can be estimated from m.
However, practically speaking a constant q case (yielding a constant slope on a T-ln(t) plot) is not observed in the late-time temperature
behavior of a geothermal well (see Figs. 7 and 8).
As discussed and validated by computational analysis below, such a method of estimating thermal conductivity seems not possible for
the typical heat transmission in a geothermal well.

Figure 9 illustrates the typical late-time wellhead temperature behaviors of two cases for a geothermal well obtained from the data given
in Tables 1 and 2 for w=4.39 and 43.9 kg/s, respectively. The wellhead temperature in Figure 9 changes according to Eq. 1. The
temperature-log(time) behaviors of both cases do not exhibit any semilog straight line relationship. The wellhead temperature for
w=4.39 kg/s case increases sharply at early times and the increase becomes lower as the time increases. For particular cases, especially
when the flow rate is high and NRa (= 2kL/(wc)) becomes small, the wellhead temperature seems to reach an apparently stabilized
value and the temperature change becomes negligibly small. Such a behavior is observed for w=43.9 kg/s case in Fig. 9.

Figure 9: Wellhead temperature versus time plot.

For the time region when the temperature change becomes negligible or assumed to be reached to an apparently stabilized wellhead
temperature, (1/q) versus log(time) plot yields a semilog straight line relationship. This can be theoretically shown when Eqs. 4 and 13
are combined to give:
10
Kutun et al.

2
= ( ) (27)

(1.7 2 )

and

1 1 1
= (1.7 2 ) + ln() (28)
2( ) 4( )

Equation 28 indicates that (1/q) versus ln(t) plot should give a semilog straight line with a slope being equal to
1
= = (29)
4( )

Figure 10 presents (1/q) versus log(t) relationship for the case with w=43.9 kg/s case. The wellhead temperature data given in Fig. 9
were used to construct the plot. The slope of the semilog straight line is determined to be 4.92x10 -4 s.m/(J-log cycle) from the (1/q)
versus log(t) plot which corresponds to 2.136x10-4 s.m/(J-ln cycle) from the (1/q) versus ln(t) plot. Assuming an average (T-Te) value of
122 oC valid for the time region in which the semilog straight line exists (t=5-100 days as observed in Fig. 10), the thermal conductivity
is calculated from the slope to be 3.05 J/m.s.oC which matches well with the input value of 2.92 J/m.s.oC.

Figure 10: (1/q) versus time plot.

The discussion above demonstrates that the wellbore heat transmission solutions at sufficiently long times converge to the cylindrical-
source solution transfering heat at constant temperature rather than the constant heat flux, and the heat transfer is controlled by
surrounding formation conditions.

One another observation deduced from the analysis of this case is that the beginning of the semilog straight line behavior is related to
the stabilization time characteristic of the wellhead temperature to be discussed in the following section of this study. Notice that the
semilog straight line behavior begins at about 5 day as seen in Fig. 10. This is about the beginning of the period in which the wellhead
temperature reaches to a stabilized and almost a constant value and the heat transmission model becomes a cylindrical source with a
constant temperature.

4. CONCLUSIONS
The geothermal wellbore heat transmission is studied here. The dimensionless time function representing the transient heat transfer
between the wellbore and the surrounding formation is assessed in detail and the adequate approximations valid for practical
engineering purposes are presented.

11
Kutun et al.

The effect of time on the wellhead temperature behavior in terms of temperature derivatives is assessed. The following conclusions are
obtained:

1) At early times, the time derivative of wellhead temperature stays at a constant value and a log-log plot of dT/dlnt versus t gives a
unit slope straight line. The early-time temperature behavior is similar to the wellbore storage phenomenon occuring in oil and gas
production wells and affecting short-time transient presure behavior.
2) The intermediate-time behavior is observed when the effect of heat transfer between the wellbore and the surrounding formation
becomes significant after the early-time convection dominated period. It is recognized when the log-log plot of dT/dlnt versus t
gives a -1/2 slope straight line and/or dT/dt versus t gives a -3/2 slope straight line.
3) The wellbore heat transmission solutions at sufficiently long times converge to the cylindrical-source solution transferring heat at
constant temperature rather than the constant heat flux, and the heat transfer is controlled by surrounding formation conditions. At
late-times when the temperature change becomes negligible or assumed to be reached to an apparently stabilized wellhead
temperature, (1/q) versus log(time) plot yields a semilog straight line relationship.

REFERENCES
Carslaw, H.S., and Jaeger, J.C.: Conduction of Heat in Solids, Oxford at the Clarendon Press, Second Edition (1959), University Press,
Oxford, p. 339.
Earlougher Jr., R.C.: Advances in Well Test Analysis, SPE Monograph Series, Vol. 5, Dallas, Texas, (1977), p.38.
Hagoort, J.: Rameys Wellbore Heat Transmission Revisited, SPE Journal, December 2004, 465-474.
Horne, R.N.: Wellbore Heat Loss in Production and Injection Wells, Journal of Petroleum Technology, January 1979, 116-118.
Kutasov, I.M.:. Dimensionless Temperature, Cumulative Heat Flow and Heat Flow Rate For a Well With a Constant Bore-Face
Temperature, Geothermics, Vol. 16, No. 5/6, (1987), 467-472.
Kutasov, I.M., Eppelbaum, L.V.: An Improved Method for Determination of Formation Temperature, Proceedings, World Geothermal
Congress 2005, Antalya, Turkey, April 24-29.
Kutasov, I.M.: Dimensionless Temperature at the Wall of an Infinite Long Cylindrical Source With a Constant Heat Flow Rate,
Geothermics, 32, (2003), 63-68.
Kutun, K., Tureyen, O.I., Satman, A.: Temperature Behavior of Geothermal Wells During Production, Injection and Shut-in Operations,
Proceedings, 39th Workshop on Geothermal Reservoir Engineering, Stanford University, Stanford, California, Febr. 24-26 2014.
Kutun, K., Tureyen, O.., Satman, A.: Revisiting the Static and Dynamic Temperature Profiles in Geothermal Wells, to be presentad at
World Geothermal Congress 2015, Melbourne, Australia, April 19-25.
Ramey, H.J.: Wellbore Heat Transmission, Journal of Petroleum Technology, April 1962, 427-435.
Ramey, H.J.: How to Calculate Heat Transmission in Hot Fluid, Petroleum Engineer, November 1964, 110-120.
Ramey, H.J.: Reservoir Engineering Assessment of Geothermal Systems, Department of Petroleum Engineering, Stanford University,
Stanford, October 1981.
Satman, A., Zolotukhin, A.B., Soliman, M.Y.: Application of the Time-Dependent Overall Heat-Transfer Coefficient Concept to Heat-
Transfer Problems in Porous Media, Society of Petroleum Engineers Journal, February 1984, 107-112.
Satman, A., Brigham, W.E.: A New Approach For Predicting the Thermal Behavior in Porous Media During Fluid Injection,
Transactions, Geothermal Resources Council, Vol. 3, September 1979.
Van Everdingen, A.F., and Hurst, W.: The Application of the Laplace Transformation to Flow Problems in Reservoirs, Trans. AIME,
Vol. 186, 305-324, (1949).

12

You might also like