Sei sulla pagina 1di 12

Building and Environment 35 (2000) 489500

www.elsevier.com/locate/buildenv

Combined simulation of airow, radiation and moisture transport


for heat release from a human body
Shuzo Murakami a,*, Shinsuke Kato a, Jie Zeng b
a

Institute of Industrial Science, University of Tokyo, Tokyo, Japan


Institute for Research in Construction, National Research Council Canada, Ottawa, Canada

Received 29 March 1999; accepted 15 June 1999

Abstract
This paper described a combined numerical simulation method of airow, thermal radiation and moisture transport for
predicting heat release from a human body. A human thermo-physiological model was also included to examine the sensible and
latent heat transfer from the human body. Flow, temperature and moisture elds were investigated with three-dimensional
Computational Fluid Dynamics (CFD). We used a low-Reynolds-number type kE turbulence model, with the generalized
curvilinear coordinate system to represent the complicated shape of the human body. The thermal radiation was calculated by
means of Gebhart's absorption factor method, and the view factors were obtained by the Monte Carlo method. We adopted
Gagge's two-node model to simulate the metabolic heat production and the thermoregulatory control processes of the human
body. The predicted results were very close to those of an actual human body in a similar situation. 7 2000 Elsevier Science
Ltd. All rights reserved.

1. Introduction
CFD technique has been greatly developed in recent
years. It is now possible for the HVAC researchers to
numerically investigate complex indoor climates with
sucient accuracy and acceptable CPU time [15].
Moreover, the coupled simulation of CFD and thermal
radiation has also been developed and widely used [6
8].1 This is of great advantage in analyzing thermal
comfort of a human body since thermal radiation
plays an important role in the thermal sensation of
humans.
* Corresponding author. Tel.: +81-33402-6231; fax: +81-337461449.
E-mail address: murakami@iis.u-tokyo.ac.jp (S. Murakami).
1
In this paper, the word ``combined simulation'' meant the interactive analysis of the relationship between the indoor climate and the
human body that included the thermo-physiological model, as shown
by Part I and II in Fig. 2. Another word ``Coupled simulation'' was
designated as the mathematical procedure for integrating CFD and
radiant heat transfer, as shown by Part II in Fig. 2.

One of the most important research targets in the


eld of HVAC is to investigate thermal sensation.
Those studies used to be carried out by means of
experiments [911]. However, recently researchers
have also begun to numerically analyze thermal sensation on the basis of coupled simulation of CFD
and radiation [3,8]. The human body not only has
a complicated physical shape but also has complex
thermo-physiological properties, thus it is very dicult to include those factors completely into the numerical simulation of an indoor climate. The
previous studies used to simplify those two aspects.
Concerning the complicated shape, most cases only
regarded the human body as a heat source without
physical shape. In a few studies, the human body
was simplied as a rectangle, a cylinder, or a ball
[12]. Among those studies, the rectangle was the
most frequently used to stand for the human body.
We understood that the presence of the physical
shape of the human body played a potential inuence on the indoor climate. Therefore the shape of
the human body should be taken into consideration

0360-1323/00/$ - see front matter 7 2000 Elsevier Science Ltd. All rights reserved.
PII: S 0 3 6 0 - 1 3 2 3 ( 9 9 ) 0 0 0 3 3 - 5

490

S. Murakami et al. / Building and Environment 35 (2000) 489500

Nomenclature
AD
Bij
cp,bl
D
Edif
Emax
Ersw
Esk
H
hfg
k
K
M
mbl
mrsw
msk
Pa
Pa,ref
Psk,s
Qcd
Qcv
Qg
Qm
Qr
Qres
Qsk
Qt
Scr
Ssk
T
Ta
Tb

skin surface area of human body, m2


Gebhart's absorption factor
specic heat of blood, 4.187 kJ/(kg K)
hydraulic diameter, m
diusion evaporative heat transfer rate from
skin, W/m2
maximum evaporative potential from skin,
W/m2
evaporative heat transfer by regulatory sweating from skin, W/m2
total evaporative heat transfer from skin, W/
m2
height of human body, m
heat of vaporization of water, 2430 kJ/kg
turbulent energy, m2/s2
eective conductance between the core node
and the skin layer, 5.28 W/(m2 K)
metabolic heat production, W/m2
blood ow rate, kg/(m2s)
regulatory sweat rate from skin, kg/(m2s)
sweat rate generated from skin, kg/(m2s)
water vapor pressure in ambient air, Pa
water vapor pressure in ambient air outside
the boundary layer of human body, Pa
saturated water vapor pressure at Tsk at skin
surface, Pa
conductive heat transfer rate, W/m2
convective heat transfer rate, W/m2
heat generation rate, W/m2
total heat transfer rate of human body, W/m2
radiant heat transfer rate, W/m2
total heat transfer rate through respiration,
W/m2
total (sensible+latent) heat transfer rate from
skin, W/m2
sensible heat transfer rate, W/m2
rate of heat storage in core node, W/m2
rate of heat storage in skin layer, W/m2
mean temperature, K (radiation),8C (the
other)
ambient air temperature,8C
mean body temperature,8C

in the balance of simulation load and simulation


precision [3,13,14]. Another simplication in the previous studies was to assume that the body surface
had uniform temperature or uniform heat production without including the thermo-regulatory processes of the body in response to the environment
[3,4,1315]. To the author's knowledge, there are no
previous numerical studies dealing with thermo-regulatory sweating [16]. Those studies apparently were

Tb,n mean body temperature at neutrality,8C


Tcr temperature of core node of human body,8C
Tcr,n neutral temperature of core node of human
body, 36.88C
Tsk temperature of skin layer,8C
Tsk,n neutral temperature of skin layer, 33.78C
wall surface temperature,8C
Tw
U
mean velocity, m/s
W
weight of human body, kg
w
skin wettedness (including diusion and regulatory sweat)
wrsw skin wettedness due to regulatory sweat
X
absolute humidity, g/kg
y
normal direction of wall, m
y + distance from wall in viscous wall unit
Greek letters
evaporative heat transfer coecient, W/(m2 k
ae
Pa)
b
actual ratio of skin mass to whole mass of
human body
ratio of skin mass to whole body mass at neubn
trality, 0.1, Ref. [11]
2 3
E
dissipation
prate 2 of turbulent energy, m /s ,
E=En(@ k/@y ) ; or surface emissivity for
radiation
l
thermal conductivity, W/(m K)
moisture conductive coecient in air, 3.082 
lx
105 kg/(m s)
Subscripts
a
air
cr
core node
i,j
cell number
in
supply opening
ref
reference point outside boundary layer of
human body
sk
skin layer
s
saturated state
n
thermal neutrality
w
wall surface
l
the rst cell near wall

not available to investigate the thermal sensation in


a warm or hot environment, because thermo-regulatory sweating must appear in that kind of situation.
Furthermore, the thermal sensation of humans is
highly dependent on the local heat transfer characteristics of the body surface. The existence and metabolic
heat production of the body can have a large inuence
on the microclimate around the body. On the other
hand, the local properties of the microclimate around

S. Murakami et al. / Building and Environment 35 (2000) 489500

Fig. 1. Floweld analyzed.

the body signicantly aect the local heat transfer


characteristics of that body. Thus it is highly important to correctly analyze the interaction (heat and
moisture transfer) between the human body and its
surrounding microclimate [16].
This study developed a combined numerical simulation method of airow, moisture transport (CFD),
and thermal radiation integrating a human thermophysiological model to reproduce the complex heat
production of the human body. This method was
applied successfully to examine the total (sensible+latent) heat transfer characteristics of the human body.
Therefore it has the potential capability of precisely
predicting the thermal sensation of humans in various
indoor environments.

491

[17]. The room was air-conditioned in order to remove


the heat and moisture production from the manikin. A
displacement ventilation system was used to achieve
the stagnant ow eld in the room space (supply temperature: 228C, supply velocity: 0.12 m/s, air ow rate:
57 m3/h, air change rate: 3.7 ACH). The walls around
the room were adiabatic for both heat and moisture,
assuming that the placement of the room would be in
the interior part of a building. The shape of the manikin was slightly simplied from that of a real human
body with feet and arms put together, close to the
body (Fig. 1), under the consideration of keeping the
balance of simulation load and precision. The manikin
had the standard height (1.651 m) and weight (65.5
kg) of a Japanese male adult, and thus the standard
surface area (1.688 m2) [3].
3. Combined numerical simulation method
The combined numerical simulation method is illustrated in Fig. 2. The simulation was composed of two

2. Floweld analyzed
The oweld analyzed is shown in Fig. 1. As stated
before, the aim of this study was to develop a combined numerical simulation system for investigating the
total (sensible+latent) heat transfer characteristics of
the human body. Our simulation target was a naked
human body (manikin) standing in a stagnant
environment.2 For the standing human body, the
metabolic heat production M was suggested as 1.7
Met (100.4 W/m2) based on the ASHRAE handbook

2
In a previous paper [3,13,14] when analyzing the heat transfer
from a human body, the internal heat transfer inside the skin surface
was not dealt with. In that case, it was not necessry to dene
whether the human body was naked or clothed. The boundary condition for that analysis was only the metabolic heat production. In
this paper, it was necessary to specify the human body as being
naked because the internal heat transfer was dealt with here. The
skin surface temperature Tsk was calculated by the interactive analysis between the skin surface and the surrounding environment. However, the eect of clothing can be modeled without diculty by
introducing the thermal and vapor resistance corresponding to the
clothing following Gagge's model into the currently developed simulation system.

Fig. 2. Flow chart of combined simulation of airow, radiation, and


moisture transport.

492

S. Murakami et al. / Building and Environment 35 (2000) 489500


Table 1
Numerical methods
Turbulence model

Low-Reynolds-number type kE model (LaunderSharma type)

Numerical schemes

Space dierence: hybrid

Grid system

Computational domain was discretized into 163,008 cells for CFD and 2232
cells for radiation with BFC. The y + of the rst cell near body and wall
surface was less than 5. Only half space was calculated due to the symmetry of
the ow eld.

parts, which were combined at body surface through


surface temperature as well as heat and mass balance
equation. The aim and actual procedure of each part
will be discussed in detail later.
The rst part was to calculate the internal heat
transfer inside the human body surface based on the
human thermo-physiological model. For simplicity, we
assumed that the heat loss through respiration was
released directly into the environment without aecting
the microclimate around the human body. According
to this policy, the heat release rate through respiration
Qres was xed as 8.7 W/m2 in advance from Eq. (A1)
(Table A1), by assuming that the body was placed in a
uniform environment (Ta=268C, Pa=1500 k Pa: the
average situation around the human body). It should
be noted that Qres was only specied as the simulation
input data due to its lesser inuence on the room environment. Other heat and mass transfer rates of the
body were predicted from this simulation. The total
(sensible+latent) heat release rate from skin Qsk was
obtained as 91.7 W/m2 (=100.48.7), by excluding
Qres from the metabolic heat production M.
The second part of the simulation was to predict the
heat and moisture transfer between the body surface
and the surrounding environment by means of a
coupled simulation of airow, radiation and moisture
transport.

3.1. Internal heat transfer: human thermo-physiological


model
Many human thermo-physiological models, including some elaborate multi-node models, were developed
during the 1970s and 1980s [911,18,19]. Among those
models, Fanger's model and Gagge's two-node model
were widely applied for evaluating thermal sensation,
because of their well-balanced performance and simplicity [17,20]. We also adopted those two models and
examined their adaptability in our numerical simulation system. By comparing the results obtained from
those two models, it became clear that Gagge's twonode model had wider adaptability, particularly in
simulating sweating [21]. Therefore this paper only
addresses the results obtained from Gagge's two-node
model. Both Fanger's model and the two-node model,
which dealt with whole-body heat balance rather than
local balance, should be applied to the analysis of the
whole body. However this study applied them locally,
i.e., to each part of the body. The issues concerning
the local application of those models will be discussed
further.
3.2. Procedure of combined numerical simulation
3.2.1. Moisture transport
Absolute humidity X was used to represent the

Table 2
Boundary conditions
Supply opening

Uin=0.12 m/s, Tin=228C, Xin=9.5 g/kg, kin=0.002U 2in , Ein=k3=2


in /(0.3D ), D = 0.2 m

Exhaust opening

U, k, E, T, X: free slip

Wall boundary

U, k, E: Uw=0, kw=(@k/@y )w=0, Ew=0


Moisture: (@X/@y )w=0 (adiabatic wall) msk=lx (@X/@y )w (body surface)
Temperature: based on heat balance equation
Qg=Qcd+Qr+Qcv
(a) adiabatic wall: Qcd=0, Qg=0
body surface: Qcd=0, Qg=Qt
(b) Qcv=l(@T/@y )w=l(TwTl)/yl
P
4
(c) Qr=sEiT 4i N
jl BijsEiT j
E: 0.98 (body), 0.95 (wall), 0.0 (opening and symmetrical surface)

S. Murakami et al. / Building and Environment 35 (2000) 489500

moisture transport. The transport equation used is


shown in Eq. (1).

 
@rX
m
rrUX r lx t rX
1
@t
sx
Here, moisture conductive coecient in air lx was
3.082  105 kg/(m s) [22]. The turbulent Schmidt
number sx was regarded as 1.0.
3.2.2. Numerical method
We omitted the other transport equations for temperature, velocity, etc, because they were in the general
form [23]. Flow, temperature and moisture elds were
calculated with three-dimensional CFD. The CFD
used the LaunderSharma type low-Reynolds-number
kE turbulence model [23]. The thermal radiation was
calculated by means of Gebhart's absorption factor
method [24], and the view factors were obtained by the
Monte Carlo method [25]. Table 1 shows the numerical methods for CFD and thermal radiation in detail.
3.2.3. Boundary conditions
Appendix A describes the derivation of the boundary conditions for human body surface based on the
two-node model. The sweat rate generated from the
skin msk was calculated by using the two-node model
in response to the room environment. This became the
boundary condition of the body surface for solving the

493

moisture transport equation (Eq. (1)). At the same


time, the sensible heat loss from skin Qt was obtained
by subtracting the latent heat release Esk from the
total heat rate of skin Qsk. The Qt was used as the
boundary condition of the human body surface for
coupled simulation of heat convection and radiation.
The method for coupled simulation of heat convection
and radiation was described in detail in the previous
study [8]. Table 2 shows the detailed boundary conditions for the human body and wall surfaces as well
as for the supply and exhaust openings.
3.2.4. Flow chart of simulation
Fig. 2 illustrates the ow chart of the combined
simulation of airow, moisture transport and radiation
based on the two-node model. For the internal heat
transfer inside the body surface [Part I], sweat rate msk
and sensible heat loss from skin Qt, was calculated by
the given skin temperature Tsk and the environmental
parameters Ta, U, X, Pa, Tw. The obtained values were
used as boundary conditions for the combined simulation of airow, radiation and moisture transport
[Part II] (i.e., interactive heat transfer between the
human body and its surrounding environment). As the
outputs of Part II, the skin temperature Tsk and the
environment parameters were renewed. The renewed
values were fed back to the internal heat transfer part
as calculation conditions. The iteration was continued
until each variation became nally convergent.

Fig. 3. Predicted velocity eld (scalar and vector, Section ABCD, m/s,).

494

S. Murakami et al. / Building and Environment 35 (2000) 489500

ow patterns near the shoulder and above the head


obtained from the visualization experiment. The predicted ow patterns agreed very well with the experimental ones. The rising stream over the head reached
a maximum velocity of 0.23 m/s from the numerical
simulation. The predicted value also agreed with the
measurement conducted by the authors, using the real
human body (0.20 m/s) and the thermal manikin
(0.21 m/s) [27]. The results were also consistent with a
previous experiment [28].
4.2. Distribution of air temperature
Fig. 5 shows the distribution of air temperature
around the manikin. A vertical temperature gradient
was formed in the room due to the displacement ventilation system. The temperature dierence between the
feet and head levels was approximately 3 or 48C.
When convection and radiation were coupled, the vertical temperature gradient was not so steep compared
to the results previously where only the convection
was dealt with [3,13,14]. That dierence was caused by
thermal radiation, which decreased the surface temperature dierence between the room walls. The ceiling, oor and other walls were assumed to be
adiabatic here. As a result, air temperature had a tendency to become uniform. In the upper region of the
room above the manikin, the air temperature was
higher and almost uniform at 278C.

Fig. 4. Flow visualization by experiment.

4. Results and discussion


4.1. Velocity distribution
The velocity distribution of Section ABCD (Fig. 1)
is shown in Fig. 3. A blanket of warm rising air was
generated around the human body, whilst the oweld
apart from the manikin was almost stagnant. When
the rising air passed from the neck to the head region,
it was accelerated due to the contraction following the
change of the body shape. In the upper region over
the manikin, a rising thermal plume was clearly
observed. The thickness of the velocity boundary layer
around the manikin surface under the shoulders
increased with respect to the height direction. This
phenomenon was similar to the ow pattern along a
vertical heated at plate [26].
The authors also carried out experiments of ow
visualization using both an experimental thermal manikin and a real human body [27]. Fig. 4 illustrated the

Fig. 5. Air temperature (Section ABCD, 8C).

S. Murakami et al. / Building and Environment 35 (2000) 489500

Fig. 6. Absolute humidity (g/kg).

4.3. Distribution of absolute humidity and relative


humidity
Figs. 6 and 7 show distribution of the absolute
humidity and the relative humidity respectively. Moisture generated from the human body went up with the
rising stream around the body surface. Consequently,
the upper part of the space had the higher absolute
humidity than the lower part. The distribution pattern
of the absolute humidity was very similar to that of
room air temperature. Concerning the relative humidity, it had higher values near the supply opening,
where the temperature was low, because relative
humidity also depends on temperature. The relative
humidity was distributed pretty uniformly from 45 to
50% through the whole room space.

Fig. 7. Relative humidity.

from the human body by convection, radiation, evaporation and respiration respectively. One of the main
purposes of this paper was to clarify the characteristics
of the heat loss from a standing human body, which
has the metabolic heat production of 1.7 Met, in a
stagnant environment. That scenario corresponded to
one of the most common situations in an actual building space. Therefore from this study the engineer can
obtain some general concept on how much heat is nor-

4.4. Wall surface temperature


Fig. 8 shows the distribution of wall surface temperature. The adiabatic walls of the room gained heat
from the human body by radiation, and then released
the same quantity of heat to the air by convection.
Thus the wall surface temperatures were generally approximately 0.38C higher than the room air temperature. The radiant heat transfer also occurred between
the walls, due to the temperature dierence between
the wall surfaces. That caused the trend of wall surface
temperature to become uniform [3,13,14].
4.5. Heat release characteristics from human body
Fig. 9 illustrates the mean heat release quantity

495

Fig. 8. Wall surface temperature (8C).

496

S. Murakami et al. / Building and Environment 35 (2000) 489500

Fig. 9. Heat release from human body (W/m2).

mally released from the human body through each


route. As shown in Fig. 9, the human body released its
metabolic heat production (1.7 Met) to the surround
environment, 49.3 W (29.1 W/m2) by convection, 65.1
W (38.3 W/m2) by radiation, 41.3 W (24.3 W/m2) by
evaporation and 14.8 W (8.7 W/m2) by respiration. As
a total percentage of the entire heat released from the
human body approximately 29.0% was by convection,
38.1% by radiation, 24.2% by evaporation and 8.7%
by respiration under the conditions of this simulation.
The human body released the most heat by radiation.
This suggested that radiation played a signicant inuence on thermal sensation of the human body.
4.6. Heat balance between manikin and room walls
Fig. 10 shows the heat balance between the manikin
and the room walls. The oor temperature was the
lowest, generally 1.08C lower than that of the other
walls (Fig. 8). Thus, the oor gained the largest radiant heat 33.8 W in total from the human body as well
as from the other walls. Conversely, the ceiling had the
highest temperature. Therefore the ceiling released heat
by radiation (5.5 W) to the other walls and gained
heat from the air by convection. The side walls of the

Fig. 11. Surface temperature of human body.

room got less heat by radiation compared to the front


and rear wall due to the smaller view factor between
the human body (main radiation heat source) and the
side walls. This can be easily understood from the location relationship since the human body faced the
front wall. Concerning the two side walls, the radiant
heat transfer rates were not equal to each other,
although their surface temperatures were almost symmetrical (Fig. 8). That slight dierence was caused by
the unsymmetrical distribution of the surface temperature of the human body. The right side of the human
body facing the supply opening had lower skin temperature than the left side (Fig. 11). Thus the side wall
with the supply opening gained slightly lower radiant
heat from the human body than the opposite side wall
with the exhaust opening.

4.7. Temperature of manikin skin surface

Fig. 10. Heat balance between human body and its environment.

Fig. 11 shows temperature distribution of the manikin skin surface. It generally ranged from 33.0 to
34.08C. It decreased below 29.08C at the right foot
facing the supply opening, and increased above 34.08C
at the neck and the shoulders. The previous experimental results [29] demonstrated a similar tendency of
temperature distribution on the skin surface of a real
human body. The mean skin surface temperature was
calculated at 33.38C. This value was slightly lower
than that of a human body in the state of physiological thermal neutrality (33.78C) with normal indoor activity [17].

S. Murakami et al. / Building and Environment 35 (2000) 489500

Fig. 12. Convective heat transfer characteristics of human body surface.

4.8. Convective heat transfer characteristics of manikin


surface
Fig. 12 illustrates the convective heat transfer rate
and heat transfer coecient of the manikin skin surface. Both distribution patterns were similar to one
another. The values were larger at the feet because of
the thinner boundary layer at that location. They
decreased with respect to the height direction of manikin. That kind of distribution corresponded to those
of the vertical heated at plate (heat and mass transfer
book). The convective heat transfer rate ranged from
20.0 to 40.0 W/m2. The convective heat transfer coecient ranged from 3.0 to 4.0 W/(m2 K). It increased to
7.0 W/(m2 K) at the feet. The mean convective heat
transfer coecient was determined to be 4.3 W/(m2
K). The results of mean value and distribution characteristics of the convective heat transfer coecient were
in agreement with previous experiments [17,30,31].
4.9. Radiant heat transfer characteristics of manikin
surface
Fig. 13(1) shows radiant heat transfer rate of the
manikin surface. Neither the temperature dierence
3
More accurately, this concept is specied as the plane radiative
temperature, which is the average at each small plane on the body
surface.

497

Fig. 13. Radiant heat transfer characteristics of human body surface.

among the room walls nor that at the manikin surface


were signicantly large, thus the radiant heat transfer
of the manikin surface ranged uniformly from 30.0 to
40.0 W/m2. The feet, however, had slightly smaller
values due to the lower manikin surface temperature
there. On average, the heat transfer by radiation (38.3
W/m2) was larger than that by convection (29.1 W/
m2).
Fig. 13(2) shows the distribution of mean radiant
temperature of the manikin surface. It was distributed
almost uniformly from 27.0 to 28.08C.3 The left and
right sides of the manikin had the low mean radiant
temperature due to the large view factors of those
parts towards the oor, where temperature was low.
4.10. Evaporative heat transfer characteristics of
manikin surface
Fig. 14(1) shows the evaporative heat transfer rate
of the body surface. It ranged from 20.0 to 30.0 W/m2.
In this study it was assumed that the total (sensible+latent) heat loss rate was the same over the whole
manikin surface. Consequently, the evaporative heat
transfer rate decreased below 20.0 W/m2 at the feet,
due to the large sensible heat transfer rate at that
point. The evaporative heat transfer rate increased
over 30.0 W/m2 at the shoulders. These results corresponded to previous experimental results [22].
Fig. 14(2) shows the evaporative heat transfer coecient of the manikin surface. The distribution characteristics were similar to that of the convective heat
transfer coecient. Due to the thin boundary layer at

498

S. Murakami et al. / Building and Environment 35 (2000) 489500

Fig. 16. Blood ow rate of human body.


Fig. 14. Evaporative heat transfer characteristics of human body surface.

the feet, a large coecient value appeared there and


decreased with respect to the height direction of the
body. The evaporative heat transfer coecient generally ranged from 60.0 to 80.0 W/(m2 k Pa) over the
manikin surface. The mean value was calculated as
70.2 W/(m2 k Pa). Using the ratio to mean convective
heat transfer coecient 4.3 W/(m2 K), the Lewis ratio
was obtained as 16.58C/k Pa, the exact same as the
value recommended by ASHRAE [17].

4.11. Skin wettedness


Fig. 15 shows the skin wettedness of the manikin
surface. The evaporative heat transfer rate at the feet
was low, causing the skin wettedness to also be low
there at 0.06. That value (0.06) suggested that only diffusive evaporation occurred and that regulatory sweating was not generated at the feet in this environmental
condition. The skin wettedness increased with respect
to the height direction of the body. It reached the
maximum value of over 0.18 at the shoulders. That
distribution property corresponded to the actual situation of a real human body [22]. The mean value was
0.11. It also corresponded to that of a human body in
thermal neutrality with the activity level of 1.7 Met.
4.12. Other physiological parameters of manikin
The core temperature of the manikin was almost
36.98C everywhere (data not shown). It was almost the
same as the neutral core temperature of 36.88C. Fig. 16
shows the skin blood ow rate of the manikin. It generally ranged from 4.0 to 6.0 g/(s m2) over the manikin
surface. The blood ow rate was below 2.0 g/(s m2) at
the feet. It increased above 7.0 g/(s m2) at the
shoulders and the neck. This distribution property corresponded to the manikin surface temperature
(Fig. 11).

5. Concluding remarks
Fig. 15. Skin wettedness of human body.

A combined numerical simulation system was devel-

S. Murakami et al. / Building and Environment 35 (2000) 489500

499

Table A1
Derivation of boundary conditions of human body surface [11]
Heat loss from respiration Qres
Qres=0.014 M(34.0Ta)+1.73  105M(5.87  103Pa)

(A1)

Heat balance equations for body core and skin layer


Scr=(QmQres)K(TcrTsk)cp,blmbl(TcrTsk)
Ssk=K(TcrTsk)+cp,blmbl(TcrTsk)Esk(Qcv+Qr)

(A2)
(A3)

Thermal neutrality for body core, skin layer, and whole body
Tcr,n=36.88C
Tsk,n=33.78C
Tb,n=bnTsk,n+(1bn)Tcr,n

(A4)
(A5)
(A6)

Thermo-regulatory control processes (vasomotor regulation and sweating)


mbl=[(6.3+200WSIGcr)/(1+0.1CSIGsk)]/3600
mrsw=4.72  105WSIGbexp(WSIGsk/10.7)
b=0.042+0.745/(3600 mbl+0.585)

(A7)
(A8)
(A9)

Temperature signals
WSIGcr=max((TcrTcr,n),0)
CSIGsk=max((Tsk,nTsk),0)
WSIGb=max((TbTb,n),0)
WSIGsk=max((TskTsk,n),0)
Tb=bTsk+(1b )Tcr

(A10)
(A11)
(A12)
(A13)
(A14)

Boundary conditions for solving moisture transport equation


Ersw=mrswhfg
wrsw=Ersw/Emax
Edif=(1wrsw)0.06Emax
w=wrsw+0.06(1wrsw)=0.06+0.94Ersw/Emax
Esk=wEmax
Emax=ae(Psk,sPa,ref)
msk=Esk/hfg

(A15)
(A16)
(A17)
(A18)
(A19)
(A20)
(A21)

Boundary conditions for coupled simulation of heat convection and radiation


Qt=QmQresEsk

(A22)

oped for predicting airow, moisture transport, and


thermal radiation based on a human thermo-physiological model. That system made it possible to investigate precisely the distributions of both sensible and
latent heat transfer characteristics at the skin surface
of the human body.
The two-node model was used here. It was originally
developed for applying to the whole body heat budget.
In its original concept it could not be applied to analyze the local heat transfer as was carried out in this
study. However, when adapting it locally to body surface, the obtained heat transfer characteristics of the
body corresponded with the actual phenomena of the
real human body. The use of the two-node model
locally at the body surface seemed tolerable in such
analyses. The multi-node model will be further examined at the next stage.
Thermal sensation indices such as SET and PMV
can be reported based on the results of this study.
Human's thermal sensation thus can be further predicted precisely by means of the combined numerical
prediction system developed in this study.

Appendix A
Boundary conditions of body surface were derived
based on the two-node model. Table A1 shows the
setup of the equations. The two-node model [10,11]
represented the body as two concentric cylinders, i.e.,
the inner cylinder referred to the body core and the
outer cylinder referred to the skin layer. The metabolic
heat production at the core was released to the environment by two routes. The major one was to transfer heat to the skin by blood ow and heat
conduction, and to be released from the skin to the environment by convection, radiation and evaporation.
The minor one was a direct release to the environment
through respiration (Eq. A1). As shown in Eqs. (A2)
and (A3), two heat balance equations for the body
core and the skin layer were built up. Here the steady
state was assumed; thus the rates of heat storage, Scr
and Ssk, were regarded as zero.
When the body was in the state of thermal neutrality
physiologically, the mean temperatures of the core
node Tcr,n, the skin layer Tsk,n, and the entire body

500

S. Murakami et al. / Building and Environment 35 (2000) 489500

Tb,n were at their neutral values (Eqs. (A4)(A6)). If


the temperature deviated from neutrality, the thermoregulatory control processes (vasomotor regulation,
sweating, and shivering) occurred. These processes
were simulated through the temperature signals as
shown in Eqs. (A7) and (A8). The temperature signals
were specied in Eqs. from (A10) to (A13). The mean
body temperature Tb was calculated from Eq. (A14).
The actual ratio of the skin mass to the whole body
mass b was calculated from Eq. (A9), because of its
dependency on the blood ow rate.
The evaporative heat loss by regulatory sweating
Ersw and the skin wettedness due to regulatory sweating wrsw were calculated from Eqs. (A15) and (A16),
respectively, by the specied regulatory sweat rate
mrsw. With no regulatory sweating, the skin wettedness
due to diusion was 0.06 for normal conditions. Then
the diusion evaporative heat loss Edif was calculated
from Eq. (A17). The total evaporative heat loss from
skin Esk was shown in Eq. (A19). The total sweat rate
generated from skin msk was simply calculated from
Eq. (A21). That value became the boundary condition
of the body surface for solving the moisture transport
equation. The sensible heat loss from skin Qt was derived as Eq. (A22). It became the boundary condition
of the human body surface for coupled simulation of
heat convection and radiation.
References
[1] Murakami S. Prediction, analysis and design for indoor climate
in large enclosures. ROOMVENT'92 1992;1:130.
[2] Kato S, Murakami S, Shoya S, Hanyu F, Zeng J. CFD analysis
of ow and temperature elds in atrium with ceiling height of
130 m. ASHRAE Transactions 1994;95(2).
[3] Murakami S, Kato S, Zeng J. Flow and temperature elds
around human body with various room air distribution, CFD
study on computational thermal manikin (Part 1). ASHRAE
Transactions 1997;103(1):315.
[4] Murakami S, Kato S, Zeng J. Numerical simulation of contaminant distribution around a modeled human body, CFD study
on computational thermal manikin (Part 2). ASHRAE
Transactions 1998;104(2).
[5] Chen Q. Computational uid dynamics for HVAC: successes
and failures. ASHRAE Transactions 1997;103(1):17887.
[6] Murakami S, Kato S, Kondo Y, Takahashi Y, Choi D.
Numerical study on thermal environment in air-conditioned
room by means of coupled simulation of convective and radiative heat transport, method for coupling convective simulation
and radiative simulation (Part 1). Transactions of the Society of
Heating, Air-Conditioning and Sanitary Engineers of Japan
(SHASE) 1995;57:10516 [in Japanese].
[7] Murakami S, Kato S, Choi D, Kobayashi H, Omori T.
Numerical study on thermal environment in air-conditioned
room by means of coupled simulation of convective and radiative heat transport, accuracy of shape factor calculation by
modied Monte Carlo method and its application for a room
with complex geometry (Part 2). SHASE Transactions
1995;59:95104 [in Japanese].
[8] Murakami S, Kato S, Zeng J. Coupled simulation of convective
and radiant heat transfer around standing human body, study

[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]

[22]
[23]
[24]
[25]
[26]

[27]

[28]
[29]
[30]
[31]

on computational thermal manikin (Part 3). Transactions of J


Archit Plann Environ Eng, AIJ 1999;515:6974 [in Japanese].
Fanger PO. Thermal comfort. Danish Technical Press, 1970.
Gagge AP, Stolwijk JAJ, Nishi Y. An eective temperature
scale based on a simple model of human physiological regulatory response. ASHRAE Transactions 1970;77(1):24762.
Gagge AP, Fobelets AP, Berglund LG. A standard predictive
index of human response to the thermal environment.
ASHRAE Transactions 1986;92(1):70931.
Brohus H, Nielsen PV. Personal exposure to contaminant
sources in a uniform velocity eld. Healthy Buildings '95, 1995.
Murakami S, Kato S, Zeng J. Development of a computational
thermal manikin, CFD analysis of thermal environment around
human body. Tsinghua-HVAC '95 1995;2:34954.
Murakami S, Kato S, Zeng J. CFD analysis of thermal environment around human body. Indoor air '96 1996;2:47984.
Kato S, Murakami S, Zeng J. Numerical analysis of contaminant distribution around a human body. ROOMVENT '96
1996;2:12936.
Kato S, Murakami S, Zeng J. Combined simulation of airow,
radiation and moisture transport for heat release from human
body. ROOMVENT '98 1998;2:14150.
ASHRAE fundamentals handbook, [Chapter 8] 1993.
Gordon RG, Rober RB, Horvath SM. IEEE Trans 1976;BME23:43444.
Stolwijk JAJ. A mathematical model of physiological temperature regulation in man. NASA contractor report, NASA CR1855, 1971.
Yokoyama S, Ochihuji S, Nagano K. Journal of the Heat
Transfer Society of Japan 1994;33(131):4150 [in Japanese].
Zeng J, Kato S, Murakami S. Coupled simulation of CFD,
radiation and moisture transport for sensible and latent heat
loss from human body, study of computational thermal manikin. SEISAN-KENKYU, Monthly Journal of Institute of
Industrial Science, University of Tokyo 1998;50(1):7885 [in
Japanese].
Nakayama S. Thermophysiology. Tokyo: Rikougakusya Press,
1987 [in Japanese].
Launder BE, Sharma BI. Application of the energy-dissipation
model of turbulence to the calculation of ow near a spinning
disc. Letters in Heat and Mass Transfer 1974;1:1318.
Gebhart B. A new method for calculating radiant exchanges.
ASHRAE Trans 1959;65:3213.
Omori T, Taniguchi H, Kudo K. Prediction method of radiant
environment in a room and its application to oor heating.
SHASE Trans 1990;42:918 [in Japanese].
Kato S, Murakami S, Yoshie R. Experimental and numerical
study on natural convection with strong density variation along
a heated vertical plate. In: Proceedings of the Ninth Symposium
on Turbulent Shear Flows, 1993. p. 512.
Hayashi T, Murakami S, Kato S, Takahashi T, Zeng J, Sakuma
K. Study on computational thermal manikin (Part 14),
Measurement of ow and temperature elds around real human
body and thermal manikin. In: Proceedings of the Annual
Technical Conference of SHASE, 1998. 2. p. 985988. [in
Japanese].
Homma H, Yakiyama M. Examination of free convection
around occupant's body caused by it's metabolic heat.
ASHRAE Trans 1988;94(1):10424.
Torikai K, Suzuki K, Tsutaki S. Thermal emission from human
body. In: Proceedings of the Second JSME-KSME Thermal
Engineering Conference, 1992.
Nielsen M, Pedersen L. Studies on the heat loss by radiation
and convection from the clothed human body. Acta Phys
Scandinav 1952;27:27294.
Rapp GM. Convective heat transfer and convective coecients
of nude man, cylinders and spheres at low velocities. ASHRAE
Trans 1973;2264:7587.

Potrebbero piacerti anche