Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

SPE-174963-MS

CFD Modeling of WBGT For an Offshore Platform


Kuochen Tsai, Shell International Exploration and Production

Copyright 2015, Society of Petroleum Engineers

This paper was prepared for presentation at the SPE Annual Technical Conference and Exhibition held in Houston, Texas, USA, 28 –30 September 2015.

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

Abstract
A wet-bulb globe temperature (WGBT) model has been incorporated into computational fluid dynamics
(CFD) software to assess heat stress for the safety of workers on an offshore platform crowded with oil
processing units. The model contains all critical elements needed to model WBGT, including conductive
and convective heat transfer from the ambient air, radiative heat from the hot processing unit surfaces,
solar load and ambient atmosphere, and the blockage of equipment, pipes and structural frames of the
platform.
The CFD model was constructed with geometry files directly imported from 3-D modeling software
used for large offshore platform structures. The imported geometry was then extracted for necessary parts
critical for the thermal modeling. Smaller features, such as deadleg pipes, nuts and bolts that won’t
influence the results are removed to reduce computational costs. The parts were then grouped according
to thermal boundary conditions to be applied, including temperature, heat transfer coefficients and
emissivity. A wrapping mesh was applied to convert the CAD geometry into a meshable domain and mesh
was generated with specified sizes that can eliminate small gaps and undesirable features ill-suited for
CFD solutions. A WBGT model was incorporated into a commercial CFD software package using an
iterative procedure from the work of LilJegren et al. (2008). The inclusion of the WBGT model greatly
enhances the predictive capability of CFD models for assessing the heat stress experienced by human
operators due to its ease of use without resorting to the modeling of human body surfaces and locations.
The model was carefully verified by comparing the CFD results with explicitly modeled WBGT sensor
surfaces and the impact of radiation, wind speed and relative humidity are shown. The modeled results for
the offshore platform show that it is necessary to insulate the processing units. The feasibility of such a
comprehensive model greatly enhances the understanding of the heat stress distribution and provides new
opportunities for remedial actions.

Introduction
With the exploration of deeper water for oil production, the oil temperature increase can be a significant
problem not only for oil processing but also for the safety of the operators. Recent data shows that some
of the offshore wells can reach 25,000 – 30,000 feet in depth with production rates exceeding 200,000
bbl/day. The temperature can reach 350 °F on the topside. Technical challenges can involve the proper
choice of elastomer seal, corrosion inhibition chemicals and other flow assurance issues. In addition, the
2 SPE-174963-MS

work environment can be extremely hot due to the approximity to congested processing units on the
platform. To ensure the work environment is safe for operations personnel and to determine the
appropriate remedial actions, understanding the heat stress load on operators during the design phase is
critical.
Wet-bulb globe temperature (WBGT) is an OSHA standard measurement for heat stress load, and a
standard portable apparatus is available for accurate measurements once a process area of interest is
operational. The measured information includes heat radiation and evaporation that can accurately
approximate the heat stress humans can experience in hot environments. Models are available to predict
WBGT, but they all require the information of heat radiation load which cannot be easily obtained due
to the difficulties in determining the correct view factors between surfaces. Computational fluid dynamics
(CFD) provides a more precise and flexible way of modeling arbitrary work environment without overly
simplified assumptions. Although more computationally expensive, the cost has dramatically lowered in
recent years and will continue to reduce in the foreseeable future. Combining CFD’s ability in accurately
computing the effect of thermal radiation and convection, and the WBGT’s ability in quantifying human
heat stress load, a 3-D map of WBGT distribution in the work area can be generated allowing assessment
of work environment safety and remedial measures in the design stage with high fidelity, and avoiding
potentially costly retrofitting activities.

Modeling WBGT
WBGT index was developed by the US Marine Corps in the 1950’s to reduce heat stress injuries in
recruits. It is a composite temperature that includes the effect of temperature, humidity, wind speed and
thermal radiation on humans and is an empirical formula. The equation states:
Eq. 1

where Tw is the natural wet-bulb temperature, Tg the globe thermometer temperature and Ta the actual
air temperature. Special devices are available to measure WBGT using a combination of three sensors,
namely, a black globe, a wetted wick and a radiation-shielded temperature sensor. The black globe is
designed to measure the influence of radiation, the wetted wick to measure the natural wet-bulb
temperature, and the sensor to measure the ambient air temperature. Although each sensor is designed with
specific thermal transfer mechanism in mind, in reality the black globe sensor measurement is influenced
both by thermal radiation and convection, and the wet-bulb sensor is influenced by a combination of
evaporation, convection and thermal radiation.
There have been numerous publications attempted to model WBGT as a predictive tool without the
actual measurement. The work by Liljegren et al. (2008) uses fundamental principles of heat and mass
transfer to model WBGT and includes all necessary heat transfer mechanisms needed for this study. Their
formulations are summarized below.
The natural wet-bulb temperature model
For the natural wet-bulb temperature, the overall energy balance equation on the wick surface accounting
for convection, evaporation and radiation states:
Eq. 2

where ⌬H is the heat of evaporation, M the molecular weight, Pr the Prandtl number, Sc the Schmidt
number, ew the saturation pressure at Tw and ea the saturation temperature of air at Ta, P is the barometric
pressure, ⌬Fnet the net radiation heat flux on the wick surface, A the wick surface area and h the
convective heat transfer coefficient of the wick. The wick is a cylinder with a diameter of 7 mm and a
length of 25.4 mm. The partial pressure of water vapor is obtained through relative humidity and the
saturation vapor pressure at temperature T calculated using the algorithm of Buck (1981, 1996). h
SPE-174963-MS 3

represents the effect of convection and can be obtained from the following formula given by Bedingfield
and Drew (1950) for a cylinder in cross flow:
Eq. 3

where Nu is the Nusselt number, Re the Reynol’s number, Pr the Prandlt number with a ⫽ 0.56, b ⫽
0.281, and c ⫽ 0.4. Solving for the convective heat transfer coefficient results in
Eq. 4

where k is the thermal conductivity of the air and D is the diameter of the cylinder. The effect of wind
speed is correlated through the Reynold number in the last term of Eq. 2. The net radiative gain by the
wick from the environment can be shown as
Eq. 5

The first term on the right is the thermal radiation emitted by the atmosphere and hot surfaces that is
absorbed by the wick, where ␴ is the Stefan-Boltzmann constant; ␧w, ␧a, and ␧sfc are the emissivity of the
wick, atmosphere and surfaces, respectively; and Ta and Tsfc are the air and surface temperature,
respectively. The second term is the radiation from the wick and the third and fourth terms are the
diffusive and direct solar irradiance, respectively, absorbed by the wick, where A1 and A2 are the areas of
the wick involved in radiation heat transfer from the solar load. The detailed calculation can be found in
the work of Liljegren et al. (2008) and is not repeated here. The calculation of A1 and A2 involves the
assumption of the view factors relative to the solar radiation from all directions, including direct, diffusive
and reflected solar irradiation.
However, in the CFD model used in this work, all the view factors can be computed using the popular
discrete ordinate (DO) model (Chowdhury, 1996) for its accuracy and efficiency in dealing with
surface-to-surface radiation. This is a finite volume implementation and can be solved simultaneously
with the convective and conductive heat transfer. In the CFD calculation, the accounting of thermal
radiation can be summed using the irradiation heat flux from all surfaces, including the sun and any hot
surfaces in the model. Thus, Eq. 5 can be reformulated to
Eq. 6

where G (w/m2) is the incident radiation flux absorbed by the wick surface A and is computed by
Eq. 7

where I is the radiation intensity solved by the DO model and ⍀ is the solid angle.
In this formulation, it is assumed that the wick absorbs and emits thermal radiation from all angles
without specifically accounting for the separate radiation sources from the sun, the ground or other hot
surfaces since it is unnecessary. The DO model calculates the radiation vectors in each computation cell
and treats both the view factors and energy sources as discretized quantities with coordinates in both
distance and view angles. The approach is more expensive than other methods since it involves the extra
dimension of view angle and which must be discretized with sufficient numbers to have good accuracy.
In general at least 2 angle discretizations in each octant are needed to reach good accuracy. The method
has been found to be in good agreement with direct surface-to-surface view factor calculations within
1–2% of global errors. However, the wick on the apparatus is not fully exposed to the ambient
4 SPE-174963-MS

environment since the bottom of the cylinder is connected to the device base and is shielded. The exposed
area is about 94% of the total surface area so a correction factor needs to be applied to both terms on the
right-hand side of Eq. 5.
The formulation for G is derived by calculating the energy balance of an infinitesimally tiny sphere in
space. In most applications the wick (7mm ⫻ 25.4 mm) and the globe (50.8 mm ID) on the measuring
device are small in comparison to the environment of interest. The deviation is expected to be insignif-
icant. The accuracy of the wet-bulb temperature is critical since it constitute 70% of the WBGT. By setting
⌬Fnet ⫽ 0, i.e., ignoring the radiation contribution, the value can be compared to the psychrometric
wet-bulb temperatures measured at night or indoors. Values obtained with this method have been found
to agree with published psychrometric charts at 101.325 kPa (sea level) and 92.6 kPa (750 m above sea
level) (Liljegren et al., 2008).

The globe temperature model


A similar analysis can be applied to the globe temperature. The formulation is much simpler since there
is no evaporation involved and only convective and radiative heat transfer are considered. The globe is
a small black sphere 2” in diameter and the energy balance can be written as
Eq. 8

The two terms on the left represents the energy lost by the globe due to radiation and convection, and
the term on the right is the energy received through radiation from the ambient air and hot surfaces as
described earlier. The value of h can be derived through the empirical correlation for a sphere as shown
below:
Eq. 9

where k is the thermal conductivity of air, and D is the diameter of the globe.

CFD Implementation
The CFD implementation is done using a commercial code. The model of an offshore platform is a small
portion of the complete platform and only the area crowded with high temperature processing units is
modeled. For the test of WBGT model accuracy, a simple model was built with geometries shown in Fig
1 where a square channel with a dimension of 6mx4mx2m (LxWxH) with both ends open to air flow in
the L dimension. Two layers of nine 2” balls are evenly placed in two heights of 0.5m and 1m. The mesh
is generated with the ball surfaces highly resolved using wall boundary layers with the near wall y⫹ value
between 1 and 10. This results in a mesh size of 5 million computational cells. The validation is done in
three ways:
1. Without wind, the balls should have temperatures equilibrated by radiation only
2. With wind and radiation to validate the model and results from CFD on the ball surfaces and the
field data as described by Eq. 8
3. With wind, radiation and relative humidity to validate the results described by Eq. 2.
SPE-174963-MS 5

Figure 1—The box and the 18 2” ball surfaces are shown on the left and a zoom-out for the temperature contour on the ball is shown
on the right. Note that the range of temperature shown is less than one degree.

Figure 1 shows the accuracy of the DO model without the influence of convection, conduction or
evaporation. All the boundaries are set to have a surface temperature of 300 °F and a 0 velocity is set the
inlet boundary on the right. It demonstrates the ability of the DO model using locally discretized angular
coordinates to capture the effect of surface-to-surface radiation. The result shows that with 4 angular
divisions and 2 pixels in each space angle division the accuracy can approach 99.9% with the averaged
ball surface temperature to be 299.5 °F. The emissivity values of all surfaces are assumed to be one for
simplicity.
In Figure 2 the wall on the far end is reduced to 90 °F to further test the accuracy of radiation
temperature calculated from incident radiation (Eq. 6) in comparison to the average surface temperature
of the small balls where the radiation effects are computed explicitly. The averaged error was found to be
less than 0.5%. This is not surprising since the ball surface is small in comparison to the box geometry,
thus the deviation from the view factor is minimal.

Figure 2—The temperature distribution at Hⴝ1m. The temperature computed based on incident radiation closely matches the average
temperature on the ball surfaces where radiation effects are computed. The averaged error for all nine balls are less than 0.5%.

Figure 3 demonstrates the importance of radiation effect on the ball surfaces. When comparing to air
temperature alone, where only conduction is considered since velocity is assumed to be zero and no
6 SPE-174963-MS

radiation interaction between the walls and air, the error is around 3– 4%. This error will be even larger
when higher velocity ambient air flow is considered.

Figure 3—Air temperature vs. ball surface temperature alone shows the importance of radiation.

For later cases, only the wall at the near end is kept at 300 °F and the rest of the wall surfaces are
reduced to 90 °F, including the inlet and outlet planes to reflect the typical temperature range for the
offshore platform of interest. Figure 4 shows the predictions of the globe temperature. It can be seen that
the empirical correlation for convective heat transfer works very well at higher temperatures and start to
deviate at lower temperatures, but is still within 1% of errors (in °K).

Figure 4 —The globe temperature contour at Hⴝ1m. The values shown here represent the combined effect of radiation, convection and
conduction without evaporation.

The validation for the evaporation model cannot be done directly using CFD since the evaporative heat
transfer on the surface needs to be modeled the same way as described by the second term on the
right-hand side of Eq. 2, which will be correct as long as the convective and radiative heat transfer are
modeled correctly. Instead the impact of relative humidity and wind speed are shown using the WBGT
formula (Eq. 1). Figure 5 shows the WBGT contours at H⫽1m with wind speed of 5.0 m/s and 1.7 m/s
from the right and a relative humidity of 80%. The higher wind speed lowers the WBGT value
significantly on the lower temperature end by more than 5 °F due to evaporation and convection.
However, the overall average WBGT is only reduced by 1 °F from 92.8 to 91.8 °F.
SPE-174963-MS 7

Figure 5—WBGT contours at Hⴝ1m with wind speed of 5 m/s (left) and 1.7 m/s (right) from the right and a relative humidity of 80%.

The impact of relative humidity can be seen clearly in Figure 6. The lower humidity allows much
higher evaporative heat transfer and lowers the average WBGT by about 8 °F and expanded the lower
WBGT zone significantly. This is consistent with the discomfort experienced in the high humidity areas
such as Houston, TX during the summer season.

Figure 6 —WBGT contours at Hⴝ1m with relative humidity of 80% (left) and 50% (right) with wind speed of 5m/s from the right.

Offshore Platform WBGT Models


In this example of an offshore platform, the heat transfer around the boarding riser skid is modeled, as the
temperature in this area is the highest in the processing area. The temperatures on the surfaces of pipes
and equipment are assumed to be known for simplicity. Two cases were studied with the hot surfaces of
the processing units specified as uninsulated (350 °F) and insulated (140 °F). The ambient air is modeled
as an ideal gas with buoyancy accounted for by turning on the gravity and modeling the density as a
function of temperature. Figure 7 shows the equipment and processing unit surfaces and the dimensions
of the simulation domain. The wind is assumed coming from the East-East-South (EES) direction at 1.7
m/s with the right-hand side pointing to the east with a temperature of 90 °F. The relative humidity is
assumed to be 80% for both cases.
8 SPE-174963-MS

Figure 7—The geometry and surfaces of processing units and pipes on a boarding riser skid. The modeled domain has a dimension
of 50’ⴛ50’ⴛ25’ (LⴛWⴛH). The red color show the hot surfaces on the processing units and pipes with oil flowing through. The wind
comes from the East-East-South direction at 1.7 m/s.

Figure 8 shows the contour of WBGT and Ta (ambient temperature) for comparison with the equipment
and pipe surface temperatures set at 350 °F, and a wind speed of 1.7 m/s from the EES direction at 90 °F.
The WBGT contour shows the strong impact from radiation, and it can be seen that the surrounding areas
to the hot surfaces is not suited for any type of work since most of the area exceeds 100 °F WBGT (Table
1). By adding insulation to make most of the equipment and pipe surface temperatures 140 °F as shown
in Figure 9, the WBGT can be lowered to below 87 °F for most of the areas. Thus, routine work can be
performed without excessive breaks.

Figure 8 —The contour of WBGT (left) and Ta (ambient temperature) (right) at 5 ft height.
SPE-174963-MS 9

Table 1—Permissible heat exposure threshold limit value (OSHA Technical Manual, Sec. III, Chap. 4)

Figure 9 —The temperature contour of the processin unit surfaces. The red-colored loactions show the highest temperature (Ta) of 350
°F at areas that cannot be insulated. The rest of the processing units and pipes surfaces are insulated to a temperature of 140 °F.

Figure 10 —The contour of WBGT (left) and Ta (ambient temperature) (right) at 5 ft height with insulated surfaces.

Conclusions
A CFD model for WBGT has been developed and demonstrated for various scenarios showing the effects
of ambient temperature, radiation, wind speed and humidity. In contrast to conventional models, the CFD
implementation enables detailed and accurate prediction of thermal radiation in 3-D geometries, thus
eliminating the need to approximate view factors on various surfaces. In a simple environment, such as
10 SPE-174963-MS

open outdoor spaces, conventional approaches can be simple and effective, but for a complex environment
such as the offshore platform, they are insufficient. The developed model combines the detailed WBGT
calculation using the heat and mass balance approaches described by Liljegren et al. (2008) and the 3-D
modeling capability of CFD with radiation. This approach eliminates the need to explicitly model the
surface of the measuring sensors and enables the WBGT to be predicted with high fidelity and accuracy
not possible before.
The effects of radiation, convection and conduction were demonstrated by carefully comparing the
results of the explicitly modeled globe sensor surfaces with the implemented WBGT field function using
the incident radiation data, ambient temperature and wind speed from the CFD model. The evaporation
is incorporated using the formula of Buck (1981). The simple channel flow model shows the importance
of radiation, wind speed and relative humidity. The offshore platform example demonstrates the com-
plexity of the work environment typical in many industries and is difficult to model by the conventional
approach.
The CFD WBGT model developed in this work enables heat stress to be analyzed with high fidelity
and accuracy for many applications in high temperature environments with complex equipment installa-
tions. The analysis can be performed in the design stage and can help to reduce costly remedial actions
later in operational life.

References
Buck, A. L. 1981. New equations for computing vapor pressure and enhancement factor. J. Appl.
Meteorol. 20: 1527–1532.
Buck, A. L. 1996. Buck Research CR-1A User’s Manual, Appendix 1.
Chowdhury B. H. 1996. Emission control alternatives for electric utility power plants. Energy Sources.
18: 393–406.
Department of health and human services (DHHS), National institute for occupational safety and
health (NIOSH). 1986. Criteria for a recommended standard – occupational exposure to hot environments.
DHHS/NIOSH Pub. No. 86 –113.
Liljegren, J. C., R. A. Carhart, P. Lawday, S. Tschopp and R. Sharp. 2008. Modeling the wet bulb
globe temperature using standard meterological measurements. J. of Occupational and Enironmental
Hygiene 5: 645–655.
Perry, R. H. and D. W. Green. 1997. Perry’s Chemical Engineers’ handbook, 7th ed.: 12-4-12-5, New
York: McGraw-Hills.

You might also like