Anda di halaman 1dari 8

Available online at www.sciencedirect.

com

Applied Thermal Engineering 28 (2008) 801808


www.elsevier.com/locate/apthermeng

Review

A survey of wind convection coecient correlations


for building envelope energy systems modeling
J.A. Palyvos *
Solar Engineering Unit, School of Chemical Engineering, National Technical University of Athens, Greece GR-15780
Received 30 September 2007; accepted 6 December 2007
Available online 15 December 2007

Abstract
The thermal losses to the ambient from a building surface or a roof mounted solar collector represent an important portion of the
overall energy balance and depend heavily on the wind induced convection. In an eort to help designers make better use of the available
correlations in the literature for the external convection coecients due to the wind, a critical discussion and a suitable tabulation is
presented, on the basis of algebraic form of the coecients and their dependence upon characteristic length and wind direction, in addition to wind speed. Finally, simple average correlations are produced from the existing ones, useful for quick, gross estimates.
2007 Elsevier Ltd. All rights reserved.
Keywords: External convective losses; Wind loss coecient; Forced convection heat loss; Review of heat convective coecients

Contents
1.
2.
3.
4.
5.
6.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
The traditional correlation for the wind loss coefficient and its variants . . . . . . . . .
Boundary layer type correlations for the wind loss coefficient . . . . . . . . . . . . . . . .
Correlations for the wind loss coefficient explicit in V, L, and/or the wind direction
External convection algorithms used in building simulation programs . . . . . . . . . .
Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction
It is well documented that the thermal losses from
external building surfaces and such solar components as
collectors, chimneys, and ventilated photovoltaic arrays,
constitute a large portion of the respective energy balance.
In support of this statement, recent calculations and midwinter data regarding a tilted solar chimney mounted on
an NTUA campus buildings rooftop showed that glaz*

Tel.: +30 210 7723297; fax: +30 210 7723298.


E-mail address: jpalyvos@chemeng.ntua.gr

1359-4311/$ - see front matter 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.applthermaleng.2007.12.005

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

801
802
804
806
806
807
807

ing/cover heat losses were in the range 26.540.0% of the


incident solar radiation, depending on duct geometry,
ambient temperature, solar energy ux, and wind velocity.
A more pronounced cover heat loss to the ambient
was recorded in the case of a ventilated photovoltaic array,
situated close to the chimney and having the same dimensions and tilt angle.
Major parameter aecting the usual modeling of losses
to the ambient from such building envelope related components is the wind convection coecient, hw, a quantity not
very well documented, improper use of which can easily
cause 2040% errors in energy demand calculations [1].

802

J.A. Palyvos / Applied Thermal Engineering 28 (2008) 801808

Nomenclature
A
C
cp
hw
j
k
L
L1, L2
LR
l
Nu
Nux
NuL
Pr
Re
Rec
Rf
St
Ta
u
v
V

surface area of plate (m2)


circumference of plate (m)
heat capacity of air (J/kg K)
forced convection coecient due to wind (W/
m2 K)
Colburn j-factor, dim/less (=St Pr2/3)
thermal conductivity of air (W/m K)
length (m)
dimensions of plate (m)
characteristic length for Re (=4A/C) (m)
length along wind direction (m)
Nusselt Number, hL/k (=St Re Pr)
local Nusselt Number
average (over L) Nusselt Number
Prandtl Number, cpl/k
Reynolds Number, LVq/l (=LV/m)
critical Reynolds Number
surface roughness factor-multiplier
Stanton Number, h/qcpVf (=Nu/RePr)
ambient temperature (K, C in Eq. (1))
wind velocity component in x-direction (m/s)
wind velocity component in y-direction (m/s)
wind speed (m/s)

Compared to radiation losses, on the other hand, external


convection losses are 34 times as big [24]. And since in
many situations there has been enough skepticism toward
the standard computational correlation for the wind loss
coecient, namely that of NusseltJurges [5], a plethora of
analogous correlations have appeared in the literature in
recent years. (One of the early reviews on the subject is that
of Cole and Sturrock [6].)
A discussion of the pros and cons of using a simple
linear correlation such as the NusseltJurges one, should be
preceded by a rough categorization of the various expressions that have been used so far. On the basis of parametric
dependence, the wind convection loss coecient for an
external at surface or blu body has appeared in the literature as:
 an experimentally determined constant value, usually
given on tables (cf. [710]);
 a very simple expression, usually linear or power law
function of the wind speed (cf. [9,1113]);
 an expression involving, in addition, a characteristic
length of the surface in question, either explicitly or
implicitly, i.e. via the Reynolds Number (cf. [1417]);
 an analogous relation which also takes into account the
wind direction with respect to the surface (cf. [1,3,18]).
On the basis of the various test rigs, whose data were
used to produce the correlations for the wind loss coecient, there are expressions stemming from:

Vaz
Vo
Vw
Vf
w
Wf
x
z
z0

wind speed at height z (m/s)


wind speed at standard conditions (m/s)
wind speed at monitored surface (m/s)
free stream wind speed (10 m above roof) (m/s)
wind velocity component in z-direction (m/s)
wind modier, dim/less (1 for windward, 0.5 for
leeward)
distance from leading edge of plate (m)
height of monitored wall above ground (m)
height at which Vf measurements are taken (m)

Greek letters
a
angle of attack ()
a, b
terrain-dependent coecients, dim/less
h
wind direction (), + east of North
l
shear viscosity of air (N s/m2)
m
kinematic viscosity of air (m2/s)
q
density of air (kg/m3)
u
yaw angle ()
Subscripts
L
average over the length L
x
local value

 wind tunnel measurements and model studies on relatively small plates and blu bodies-obstacles (cf. [19
21]);
 full-scale/eld data recordings, i.e. measurements on
actual building facades and roofs (cf. [2226]).
Thus, the prospective designer/modeler must be aware of
the diversity of the available correlations and must make
sure that he has examined the specic conditions under
which they have been produced, before he can safely use
them. After all, there is an obvious lack of physical equivalence between easily controlled indoor experimental studies
and the hard reality of the eld. Out there, monitoring the
wind or establishing uniform conditions for the relevant
measurements is a formidable job. It has been reported,
for example, that average wind speed measurements in the
proximity of an external surface such as a solar collector
cannot have less than a 0.5 m/s variation [27].
2. The traditional correlation for the wind loss coecient and
its variants
In its most general form, the NusseltJurges correlation
between the wind convection coecient, hw, and the parallel to the surface component of the wind velocity, Vw,
which drives the convection can be written as


 
n 
294:26
hw 5:678 a b
V w 0:3048
1
273:16 T a

J.A. Palyvos / Applied Thermal Engineering 28 (2008) 801808


Table 1
Values for the constants in Eq. (1) [28,29]
Vw < 4.88 m/s
Surface texture
Smooth
Rough

a
0.99
1.09

Table 3
Relation of the parallel component of the wind velocity to the free stream
value [22]

4.88 6 Vw < 30.48 m/s

0.21
0.23

1
1

0
0

0.50
0.53

n
0.78
0.78

where a, b, n are empirical constants, and Ta the ambient


temperature in C. The correction in the innermost parenthesis is dictated by the fact that the original correlation
was developed for a reference temperature of 294.26 K
(21.1 C) [28,29]. (This temperature correction is equivalent
to a density correction, which is necessary since the mass
velocity rather than the linear velocity is more appropriate
for the description of forced convection [30]). The three
constants in Eq. (1), which take values that depend on
the external surface texture and the wind speed, are listed
in Table 1, giving hw in SI units.
Actually, the original 1922 global correlation of NusseltJurges [5] based on their copper plate data, if written
for SI units would have the form
0:6V w
hw 7:13V 0:78
w 5:35e

that is, it includes a decay term. This equation was specialized two years later by Jurges [31] for three types of surfaces (cf. Table 2). However, it was the original authors
who rst suggested that, for practical calculations, it is sufcient to use a linear interpolation formula for wind speeds
up to 5 m/s and to ignore the second term for higher speeds
[5]. In SI units, the proposed original linear equation would
be
hw 5:8 3:95V w

having constants which are very close to the values used in


recent years on the basis of the McAdams [28] recast of Eq.
(1), which for smooth surfaces is
hw 5:7 3:8V w

This last and much quoted correlation has been widely


used in modeling, simulations, and relevant calculations
(cf. [3234]), in spite of its shortcomings. It has been argued, for example, that this dimensional equation includes
radiation loss in addition to convection [35], and that the
average-across the surface-wind speed as well as its direction must be considered [36]. Moreover, the data which
Table 2
Values for the constants in Eq. (1) based on the original correlations of
Jurges [31], ignoring the decay terma
Vw < 4.88 m/s

4.88 6 Vw < 30.48 m/s

Surface texture

Hydraulically smooth
Rolled
Very rough

0.973
1.005
1.087

0.214
0.214
0.226

1
1
1

0
0
0

0.499
0.497
0.522

0.775
0.780
0.784

0.6 V

803

The decay term is ae


, where a takes the values 5.12, 5.35, and 5.84
for the three surface types, respectively.

Vf > 2 m/s
Vf < 2 m/s

Windward

Leeward

Vw = 0.25Vf
Vw = 0.5

Vw = 0.3 + 0.05Vf

generated the above equations were taken on a vertical


square copper plate with 0.5 m sides subjected to parallel
ow of air, a situation that is hard to meet in real life.
In order to circumvent the lack of proper values for Vw,
Ito et al. [22] tried to correlate this parallel to the surface
component with the free stream velocity, Vf, introducing,
at the same time, gross wind direction, i.e. dierentiating the results for windward1 and leeward conditions (cf.
Table 3). This procedure, which at the time was also
adopted by ASHRAE [37], allows the use of a single correlation for the convection component of the outside heat
transfer coecient for either direction, namely
hw 18:63V 0:605
w

This power law equation (which, aside from the S.I. units,
was actually derived by Kimura on the basis of earlier data
of his Japanese group [37, p. 78]), as well as the linear Eqs.
(3) and (4) have been adopted as algebraic forms by most
researchers for their own data tting procedures. An extensive but not exhaustive tabulation of such equations are given in Tables 4 and 5.
It turned out that in many cases the linear regression
equations were equally eective in tting the experimental
data, even though fundamental heat transfer theory predicts a power relation between convective coecient and
wind speed. On the basis of data generated by thirty such
linear correlations listed in Table 4 for windward surfaces,
a purely empirical average correlation of this simple but
convenient form can be derived, namely
hw 7:4 4:0V f

windward

In the wind velocity range 04.5 m/s, the maximum deviations of the individual windward equations predictions
from those of Eq. (6) average to 18%. A similar average
correlation can be derived for leeward surfaces, using the
remaining six correlations of Table 4
hw 4:2 3:5V f

leeward

In the same wind velocity range, the maximum deviations


of the individual leeward equations predictions from those
of Eq. (7) average to 22%.
It should be noted at this point that many of the individual correlations listed in Tables 4 and 5 use constants with

A surface or data are classied as windward if the angle of incidence


between the normal to the monitored surface and the wind direction is less
than 90 and leeward for all other directions [23].

804

J.A. Palyvos / Applied Thermal Engineering 28 (2008) 801808

Table 4
Linear equation form for the wind convection coecient (W/m2 K): hw = a + bV
a

Comments

References

5.8
5.7
6.2
6.05
5.82
7.82
8.9
10.7
23
11.4
5.7
0
5.8
8.7
2.8
6.22
6.22
7.55
5.8
4.5
8.55
0.036
5.1
5.1
7.0
6.47
4.955
8.91
4.93
10.03
12.2
8.3
6.5
6.42
4.47
4.214
5.82
5.5

3.95
3.8
4.3
4.08
4.02
3.50
3.71
4.96
5.7
5.7
6.0
5.7
2.9
9.4
3.0
1.824
0.4864
4.35
4.1
2.9
2.56
2.2
1.7
1.7
2.1
6.806
1.444
2.00
1.77
4.687
6.548
2.2
3.3
3.96
10.21
3.575
4.07
2.2

Wind tunnel measurement (WTM), plate, parallel ow, Vw < 5 m/s


WTM, plate, parallel ow, Vw < 5 m/s, smooth surfaceb
WTM, plate, parallel ow, Vw < 5 m/s, rough surface
WTM, vertical plate, rough surface, V 6 5 m/s
WTM, vertical plate, rolled surface, V 6 5 m/s
Very smooth surface, no speed limit
Smooth wood, plaster, no speed limit
Cast concrete and smooth brick, no speed limit
WTM, exposed face of a 23 cm cube for the range 310 m/s
Nocturnal eld measurements (NFM) on heated strips, exposed surface
NFM on heated strips, normal surface
NFM on heated strips, leeward facing surfaces
NFM, wall, Vf > 3 m/s, windward (if leeward and Vf > 4 m/s, hw = 13 W/m2 K)
NFM, Vw > 4 m/s, leeward, h independent of wind direction
Vw, revised Ref. [5] data to exclude radiation and free convection contribution
Kimuras 4th oor model, eld measurement (FM) on window, windward surface
Kimuras 4th oor model, FM on window, leeward surface
NFM in a Canadian Arctic location, window, Vf rooftop speed
Based on Ref. [5] data, Vw (for roofs design Vs are 1, 3, and 9 m/s)
Smooth surfaces (glass, paint) at ordinary temperatures (rough: 50% higher)
Rectangular plate exposed to varying wind directions, Vbar sqrt(u2 + v2 + w2)
Laboratory measurements, inclined and yawed real collector, Vw, leeward
FM on facade of tall building, Vw = f1(Vf) = 1.8Vf + 0.2 windward
FM on facade of tall building, Vw = f2(Vf) = 0.2Vf + 1.7 leeward
Flat plate PV module, experimental, 1.0 < Vw < 1.5 m/s, 0 < Ta < 35
ASHRAE/DOE-2 model, rough surfacesa excluding radiation
Daytime FM on central region of vertical wall, Vf, plate shielded from sun
FM on plane, smooth facade test surface, Vf (Vw = 0.68Vf  0.5), windward
FM on plane, smooth facade test surface, Vf (Vw = 0.68Vf  0.5), leeward
Indoor laboratory measurements on box-type solar cooker
Indoor laboratory measurements on basin-type solar still
FM, collector mimic on 35 pitched roof, Vf, incidence angle i = 0
FM, collector mimic on 35 pitched roof, Vf, incidence angle i = 90
Multipoint FM, V = sqrt(avg(u2 + v2 + w2)), developed turbulent boundary layer on horizontal roof
Multipoint FM, V = sqrt(avg(u2 + v2 + w2)), on vertical wall
Collector glass cover
Based on photovoltaic systems analysis
Outdoor measurements on PV modules without considering wind direction

[5]
[28]
[4]
[30]
[30]
[39]
[39]
[39]
[6]
[6]
[40]
[6]
[22]
[22]
[35]
[41]
[41]
[42]
[43]
[11]
[44]
[3]
[23]
[23]
[65]
[45]
[13]
[24]
[24]
[12]
[12]
[46]
[46]
[26]
[26]
[9]
[67]
[68]

a
For smooth surfaces hw = 3.12 + 3.83V  0.047 V2 excluding the constant 5.11 W/m2 K radiation contribution. A recent similar equation is
hw = 12.667 + 1.5946V + 0.0041V2 [63].
b
ASHRAE proposes a = 5.62, b = 3.9 ([60], p. 3.14). Another variant is a = 5.67, b = 3.86 [64].

more signicant gures than one would expect to obtain


during a wind speed measurement in the eld. Also, small
dierences in the values of the constants in the NusseltJurges equations, as reported by various authors, could be the
result of units-conversion errors which probably have been
propagated over the years by indirect referencing.

j-factor, expressing the analogy between (sublimation)


mass transfer and (convection) heat transfer. Using the relNu
, where
evant denitions, j = St Pr1/3, and St qcphV f RePr
St is the Stanton Number, Pr the Prandtl Number, q the
density, and cp the heat capacity of air, respectively, the
average Nusselt number, NuL , on an inclined rectangular
plate subject to an oncoming airow can be written as

3. Boundary layer type correlations for the wind loss


coecient

NuL 0:86Re1=2 Pr1=3

Thermal boundary layer theory has led to correlations


for hw or, equivalently, for the Nusselt Number, Nu, which
implicitly involve a characteristic length via the Reynolds
Number, Re, which also appears in the correlations. In a
series of wind tunnel experiments on naphthalene plates,
Sparrow and his group [18,19,15,38] have produced correlations of the form j = aReb, where j is the familiar Colburn

20:000 < Re < 90:000

In this correlation, and in similar ones involving square


plates or plates with stabilizing extensions (cf. Table 6),
the characteristic length for Re is LR = 4A/C, where A is
the plate area and C its circumference. If the plate is rectangular, with sides L1 and L2, then
LR

2L1 L2
L1 L2

J.A. Palyvos / Applied Thermal Engineering 28 (2008) 801808

805

Table 5
Power law form for the wind convection coecient (W/m2 K): hw = a + bVn
a

0
0
0
0
0
0
0
0
0.685
0
0
0
0
0
0
0
0
18.192

7.13
7.11
7.52
7.2
7.6
6.97
6.60
18.65
11.8
2.38
2.86
16.15
16.25
16.21
14.82
15.06
9.4
0.0378Tav

0.78
0.775
0.784
0.78
0.78
0.666
0.6
0.605
0.5
0.89
0.617
0.397
0.503
0.452
0.42
0.53
0.5
0.8

Comments

References
a

Wind tunnel measurement (WTM), plate, parallel ow, 5 < Vw < 24 m/s
WTM, plate, parallel ow, hydraulically smooth surface, 5 < Vw < 24 m/s
WTM, plate, parallel ow, very rough surface, 5 < Vw < 24 m/s
Recast of data in [5,31], V > 5 m/s, smooth surface
Recast of data in [5,31], V > 5 m/s, rough surface
WTM, at plate, measurement on the rear
Vertical surface behind wedge-separated subsonic ow
Field measurements (FM), VwKimuras 6th oor modelb
WTM, small (0.16 m) collector mimic
FM, window, low-rise building, forced convection only, windward (MoWiTT)c
FM, window, low-rise building, forced convection only, leeward (MoWiTT)c
FM, at vertical panel, windward, Vw = f(Vf) = 0.68Vf  0.5 and 0.2Vf  0.1
FM, at vertical panel, Leeward, Vw = f(Vf) = 0.157Vf  0.027
FM avg correlation for combined windward and leeward conditions
FM on a 6th oor vertical surface in 200 mm recess, windwardd
FM on a 6th oor vertical surface in 200 mm recess, leewardd
FM, collector mimic on 35 pitched roof, Vf
External coecient in large commercial ducts, Tav = (Tduct,surf + Texterior)/2

[5]
[31]
[31]
[28]
[28]
[47]
[48]
[37]
[49]
[45]
[45]
[24]
[24]
[24]
[50]
[50]
[46]
[66]

Note. Ref. [69] examines convective cooling of photovoltaics.


a
ASHRAE proposes b = 7.2 for 5 6 Vw 6 30 ([60], p. 3.14).
b
Vw = 0.25Vf for Vf > 2 m/s, Vw = 0.5 for Vf 6 2 m/s (windward) and Vw = 0.3 + 0.05Vf (leeward) [45].
c
Same for DOE-2 with an additional multiplier for rough surfaces, Rf.
d
Also: correlations for 4 more depths and alternate correlations involving Vf instead of Vw.
Table 6
Boundary layer equation form for the wind convection coecient (W/m2K): Nu = a RebPrc + d
a

Comments

References

0.10
0.20
0.42
0.931
0.86
0.930
0.86
0.0253
0.036
0.032
1.23
0.90
0.568
1.067
f2(Re, Pr)c
0.023

0.666
0.666
0.6
0.5
0.5
0.5
0.5
0.8
0.8
0.8
0.5
0.5
0.524
0.466
0.8
0.891

0
0
0
0.333
0.333
0.333
0.333
0
0.333
0
0.333
0.333
0
0
1
0

0
0
0
0
0
0
0
3
f1(Pr)b
84.5
0
0
0
0
0
0

[47]
[47]
[14]
[18]
[19]
[38]
[15]
[20]
[3]
[3]
[51]
[51]
[52]
[52]
[9]
[26]

0.0296Rf
0.037Rf
0.664Rf
0.037Rf

0.8
0.8
0.5
0.8

0.333
0.333
0.333
0.333

0
0
0
f3(Pr)d

Wind tunnel measurement (WTM), vertical plate, windward


WTM, vertical plate, leeward
WTM, considers house as equivalent sphere
WTM, global correlation for inclined (attack) and yawed square plate
WTM, global correlation for inclined rectangular plate of nite width
WTM, global correlation for pitch and yaw-square plate
WTM, global correlation for collector with extension surfacesa
WTM, local Nu equation for mixed (lam. and turbulent) ow over a at plate
Laboratory, avg Nu on inclined and yawed real collector, with L = L(L1, L2, u)e
Laboratory, avg Nu on inclined and yawed real collector (Pr = 0.706)
WTM, plate ush on wooden roof model, angle of attack a < 40
WTM, plate ush on wooden roof model, angle of attack a P 40
WTM, at-plate model collector ush on 30 roof of model house
WTM, at-plate model collector ush on 45 roof of model house
Nu for parallel to plate ow, turbulent boundary layer
Multipoint eld measurement (FM), Re = (l/m)sqrt(avg(u2 + v2 + w2)), developed turbulent
boundary layer on horizontal roof
FM, local Nu for horizontal roof, turbulent ow Rf surface roughness multiplier
FM, avg Nu, horizontal strip of length L, turbulent ow, DT = 0, L > xc  0
FM, avg Nu, horizontal strip of length L, turbulent ow, DT < 0, L > xc = 5  105l/(qVf)
FM, avg Nu, horizontal strip of length L, turbulent ow, DT < 0, L P xc = 5  105l/(qVf)

a
b
c
d
e

[25]
[25]
[25]
[25]

The hydrodynamic dimensions are used for the characteristic length in Re.
f1(Pr) = 95.0Pr1/3.
f2(Re, Pr) = 0.037/(1 + 2.443Re0.1(Pr2/3  1).
f3(Pr) = 871RfPr1/3.
The characteristic length is L = L1L2/(L1 cos u + L2 sin u), with L1, L2 the collector side lengths and u the yaw angle.

The choice of such a length was dictated by the intuitive


involvement of the surface area of the plate, and the very
high transfer rates observed at the edges, thus bringing also
the circumference into the picture [19]. Eq. (9) is nothing
more than the simplest combination which yields a length
dimension.

Another experimental series involving a real (albeit halfsize) solar collector under controlled environmental conditions in the laboratory [3], produced a slightly dierent correlation for the average Nusselt Number, namely,
4=5

NuL 0:036ReL Pr1=3  95Pr1=3

10

806

J.A. Palyvos / Applied Thermal Engineering 28 (2008) 801808

The critical Reynolds Number was estimated at 70,000,


implying an early development of turbulence which can
be explained by the obvious lack of a sharp leading edge
on the collector. As the latter was both inclined and yawed
with respect to the on-coming air, the appropriate characteristic length for Re was
L

L1 L2
L1 cos / L2 sin /

u < 90

11

with a similar equation for yaw angles, u, greater than 90


[3]. It turned out, however, that, as in the case of Sparrows
yawed plate [18], the inuence of the yaw angle was small,
over the entire range of us tried (0, 90, 135, and 180), but
not insignicant. The relevant reduction of hw with u was
in the range 515% [3]. On the other hand, the wind direction has also been found not to have a signicant eect in
the case of large walls [13].
Table 6 includes additional boundary layer type correlations. Among them, is the following equation for the local
Nusselt Number, Nux,
Nux 0:037Rf Re4=5 Pr1=3  871Rf Pr1=3

12

in which x denotes distance from the leading edge and Rf is


a surface roughness dependent convection multiplier,
assuming values in the range 1.00 < Rf < 2.10. (The similarity between Eqs. (10) and (12) is worth noting.) The same
authors also give relations involving explicitly the wind
incidence angle, h, for the forced convection heat transfer
coecient averaged over rectangular roof strips [25].
4. Correlations for the wind loss coecient explicit in V, L,
and/or the wind direction
A number of correlations for hw, explicit in V and L are
also listed in Table 7. Some of them have been the result of

evaluation of the thermo-physical properties which participate in the Reynolds and Prandtl Numbers, for specic
temperatures, although the latter are not always reported
(cf. [17]). They have the general form hw = aVb/Lc, that
is, they include a decay of the wind convection coecient
along the surface in the direction of the wind. For example,
the fully turbulent ow convective coecient can be written as
hw 5:74V 4=5 L1=5

13

for a at surface such as a solar collector [17]. If the latter is


ush-mounted on the (inclined) roof of a house, then


8:6V 0:6
hw max 5; 0:4
14
L
where L is the cube root of the house volume in meters [32,
p. 166]. The constant 5 in this last relation represents the
minimal hw value which is observed in solar collectors under zero wind.
The rest of the correlations in Table 7 are explicit only in
V, having the form hw = aVb, i.e. without the decay factor.
Each one is given for either wall surfaces or roofs of lowrise isolated buildings, and for a particular wind direction.
As an example, for a 45 surface-to-wind angle the relevant equation for walls is
hw 3:34V 0:84
f

15

according to recent CFD calculations [1].


5. External convection algorithms used in building simulation
programs
In view of the lack of a universally acceptable wind convection coecient or correlation (as clearly demonstrated
by the large, yet non-exhaustive, compilations of Tables

Table 7
Explicit in V, L form for the wind convection coecient (W/m2 K): hw = aVbLc
a

Comments

References

8.6
5.79
2.537WfRfa
2.537WfRfa
5.74
5.1
11.42
5.15
3.34
4.78
4.05
3.54
5.11
4.60
3.67

0.6
0.8
0.5
0.5
0.8
0.5
0.891
0.81
0.84
0.71
0.77
0.76
0.78
0.79
0.85

0.4
0.2
0.5
0.5
0.2
0.5
0.109
0
0
0
0
0
0
0
0

Convection over buildings, L = cube root of building volume


Single at-plate PV panel, Vw P 0.3 m/s, L = geometric scale
BLAST model: V = Vaz, wind speed modied for height zb
TARP model:V = Vaz, wind speed modied for height zc
Fully turbulent ow over horizontal, constant temperature surface
Field measurements (FM), plates, global, 313 K mean plate-air temperature
FM, local h, V = sqrt(avg(u2 + v2 + w2)), both for horizontal and vertical enveloped
FM, walls of isolated, low-rise building, 0 angle of attacke
FM, walls of isolated, low-rise building, 45 angle of attacke
FM, walls of isolated, low-rise building, 90 angle of attacke
FM, walls of isolated, low-rise building, 135 angle of attacke
FM, walls of isolated, low-rise building, 180 angle of attacke
FM, roof of isolated, low-rise building, 0 angle of attacke
FM, roof of isolated, low-rise building, 45 angle of attacke
FM, roof of isolated, low-rise building, 90 angle of attacke

[14]
[65]
[16]
[16]
[17]
[46]
[26]
[1]
[1]
[1]
[1]
[1]
[1]
[1]
[1]

a
b
c
d
e

Wf = wind modier (1 for windward surfaces, 0.5 for leeward ones), Rf = surface roughness multiplier, L = (surface area/perimeter).
Vaz = V0(z/z0)1/a (z0 = 9.14, a is a terrain-dependent coecient).
Vaz = V0b(z/z0)a (a, b are terrain-dependent coecients).
Pr has been evaluated at 293 K.
DT = surface-to-air temperature dierence = 10 K, wind speed 15 m/s.

J.A. Palyvos / Applied Thermal Engineering 28 (2008) 801808

47), users of large building energy simulation programs


such as EnergyPlus [53], TRNSYS [54], and ESP-r [55],
should be able to choose the most suitable form oered
by the respective program. In fact, they should be also
given the option to supply their own model, as the correlations for the external convection coecient adopted
by such programs are not standardized [56].
EnergyPlus, for example, oers six outside convection
algorithms: a simple second degree polynomial in Vw proposed by ASHRAE [8] (which, however, includes a constant radiation component of about 5 W/m2 K), the more
detailed algorithm of Sparrow et al. [19] as well as similar
expressions in the BLAST and TARP programs [16] and,
nally, the MoWiTT and the DOE-2 models [45]. The
MoWiTT algorithm oers a reasonable balance between
accuracy and ease of use [57] while the DOE-2 model is a
combination of the MoWiTT and BLAST algorithms
(see Tables 4, 5 and 7 for the respective expressions).
Most of the relevant components in TRNSYS, on the
other hand, ask the user to supply the heat transfer coecient (although the linear form of Eq. (4) is built in, for
example, in the Type 19 subroutine). The most popular
algorithm used for the external convective coecient is
the quadratic polynomial in Vw [58,59], whose constant
coecients are tabulated for six types of surface texture
[60].
Finally, external convection in ESP-r is modeled via
time varying wind driven convection coecients, calculated
with a choice of two dierent correlations: a simple linear
one (cf. [35]), with suitable amendment of the velocity value
on the basis of surface orientation and wind direction relative to the surface [61], and the MoWiTT model [45]. (The
ESP-r repertoire, however, is much more exible when it
comes to internal building surfaces, for which it oers an
option of ve empirical correlations and xed values for
the internal convection coecient [62]).
6. Conclusions
Since the importance of the external convection coecient due to the wind in, practically, all the energy calculations involving the building envelope has been well
documented, the prospective designer/modeler must be
extra careful when seeking a correlation for hw to suit
his/her needs. This is also true when one uses the major
building energy analysis programs, which accept user supplied correlations and constant values for hw or oer a
menu of such correlations.
Being aware of the diversity of the available expressions,
some of which are the result of laboratory tests while others
stem from actual eld measurements, one must make sure
that he/she has fully examined the specic conditions under
which they have been produced, before he/she can safely
use any one of them. Moreover, the understandable lack
of physical equivalence between certain experimental studies and the reality of the eld, should help in narrowing
down the options. If, on the other hand, one merely seeks

807

a quick gross estimate, one can always use the average


simple correlations proposed in this work, i.e. Eqs. (6)
and (7).
In view of the above, the expressions from the literature
which are listed in Tables 47 should prove useful, provided the reader has access to the original works cited
therein in order to assess the respective applicability to
his/her own problem at hand. However, the obvious lack
of generality of the existing wind convection coecient correlations presents a challenge for future research. More
realism is needed, i.e. eld rather than laboratory measurements, as well as some sort of standardization in the choice
of such things as the wind velocity sensors or the measurement topology, e.g. height above ground and/or distance
from the facade wall or roof, etc. In this way, a much smaller set of well proven and generally accepted correlations
may emerge, which will greatly help the designer/modeler.
References
[1] M.G. Emmel, M.O. Abadie, N. Mendes, New external convective
heat transfer coecient correlations for isolated low-rise buildings,
Energ. Build. 39 (2007) 335342.
[2] P.I. Cooper, The eect of inclination on the heat loss from at-plate
solar collectors, Solar Energy 27 (5) (1981) 413420.
[3] G. Thomaidis, J.A. Palyvos, N.G. Koumoutsos, The inuence of the
wind on the eciency of the at-plate solar collector, in: Proceedings
of the 1st National Conference on Soft Energies, Institute of Solar
Engineering, Salonica, Greece, 2022 October, 1982, HEX, pp.115
(in Greek).
[4] M.G. Davies, Build. Heat Transfer, John Wiley & Sons Ltd.,
Chichester, England, 2004, 5.7.
[5] W. Nusselt, W. Jurges, Die Kuhlung einer ebenen Wand durch einen
Luftstrom (The cooling of a plane wall by an air ow), Gesundheits
Ingenieur 52. Heft, 45. Jahrgang, (30 Dezember 1922) pp. 641642.
[6] R.J. Cole, N.S. Sturrock, The convective heat exchange at the
external surface of buildings, Build. Environ. 12 (4) (1977) 207214.
[7] ASHRAE, Cooling and heating load calculation manual, GRP 158,
ASHRAE, New York, second printing 1979, p. 3.12.
[8] ASHRAE, Handbook of Fundamentals, Atlanta, GA, USA, 1993.
[9] U. Eicker, Solar Technologies for Buildings, John Wiley & Sons Ltd.,
Chichester, England, 2003, 3.1.10.4.
[10] P. Konttinen, T. Carlsson, P. Lund, T. Lehtinen, Estimating thermal
stress in BIPV modules, Int. J. Energ. Res. 30 (2006) 12641277.
[11] P.J. Lunde, Solar Thermal Engineering, John Wiley & Sons, New
York, USA, 1980, p. 17.
[12] S. Kumar, V.B. Sharma, T.C. Kandpal, S.C. Mullick, Wind induced
heat losses from outer cover of solar collectors, Renew. Energ. 10 (4)
(1997) 613616.
[13] S.E.G. Jayamaha, N.E. Wijeysundera, S.K. Chou, Measurement of
the heat transfer coecient for walls, Build. Environ. 31 (5) (1996)
399407.
[14] J.W. Mitchell, Heat transfer from spheres and other animal forms,
Biophys. J. 16 (1976) 561.
[15] E.M. Sparrow, S.C. Lau, Eect of adiabatic co-planar extension
surfaces on wind-related solar-collector heat transfer coecients,
Trans. ASME J. Heat Transfer 103 (1981) 268271.
[16] T.M. McClellan, C.O. Pedersen, ASHRAE Trans. 103 (2) (1997) 469
484.
[17] E. Sartori, Convection coecient equations for forced air ow over
at surfaces, Solar Energy 80 (2006) 10631071.
[18] E.M. Sparrow, K.K. Tien, Forced convection heat transfer at an
inclined and yawed square plate application to solar collectors,
Trans. ASME J. Heat Transfer 99 (1977) 507512.

808

J.A. Palyvos / Applied Thermal Engineering 28 (2008) 801808

[19] E.M. Sparrow, J.W. Ramsey, E.A. Mass, Eect of nite width on
heat transfer and uid ow about an inclined rectangular plate,
Trans. ASME J. Heat Transfer 101 (1979) 199204.
[20] X.A. Wang, An experimental study of mixed, forced, and free
convection heat transfer from a horizontal at plate to air, Trans.
ASME J. Heat Transfer 104 (1982) 139144.
[21] J.L. Francey, J. Papaioannou, Wind-related heat losses of a at-plate
collector, Solar Energy 35 (1985) 1519.
[22] N. Ito, K. Kimura, J. Oka, A eld experiment study on the convective
heat transfer coecient on exterior surface of a building, ASHRAE
Trans. 78 (1972) 184191.
[23] S. Sharples, Full-scale measurements of convective energy losses from
exterior building surfaces, Build. Environ. 19 (1984) 3139.
[24] D.L. Loveday, A.H. Taki, Convective heat transfer coecients at a
plane surface on a full-scale building facade, Int. J. Heat Mass
Transfer 39 (8) (1996) 17291742.
[25] R.D. Clear, L. Gartland, F.C. Winkelmann, An empirical correlation
for the outside convective air-lm coecient for horizontal roofs,
Energ. Build. 35 (2003) 797811.
[26] A. Hagishima, J. Tanimoto, Field measurements for estimating the
convective heat transfer coecient at building surfaces, Build.
Environ. 38 (7) (2003) 873881.
[27] G.L. Morrison, D. Gilliaert, Unglazed solar collector performance
characteristics, Trans. ASME J. Solar Energy Eng. 114 (1992) 194200.
[28] W.H. McAdams, Heat Transmission, third ed., McGraw-Hill
Kogakusha, Tokyo, Japan, 1954, p. 249.
[29] J.A. Clarke, Energy Simulation in Building Design, second ed.,
Butterworth-Heinemann, Oxford, England, 2001, 7.6.2.
[30] A. Schaak, Industrial Heat Transfer, Chapman & Hall, London,
1965.
[31] W. Jurges, Der Warmeubergang an einer ebenen Wand (Heat transfer
at a plane wall), Beiheft Nr. 19 zum Gesundh.- Ing. (1924),
appearing in Gesundheits-Ingenieur 9. Heft, 48. Jahrg., 1925, p. 105.
[32] J.A. Due, W.A. Beckman, Solar Energy Thermal Processes, third
ed., John Wiley & Sons Ltd., Hoboken, NJ, USA, 2006, 3.15.
[33] D.Y. Goswami, F. Kreith, J.F. Kreider, Principles of Solar Engineering, second ed., Taylor and Francis, Philadelphia, PA, USA,
2000, p. 98.
[34] G.N. Tiwari, Solar Energy-Fundamentals, Design, Modeling and
Applications, Alpha Science International Ltd., Pangbourne, UK,
2002, p. 74.
[35] J.H. Watmu, W.W.S. Charters, D. Proctor, Solar and wind induced
external coecients for solar collectors, Comples. Int. Rev. dHellio
Tech. 2 (1977) 56.
[36] M.V. Oliphant, Measurement of wind speed distributions across a
solar collector, Solar Energy 24 (1980) 403405.
[37] ASHRAE Task Group, Procedure for determining heating and cooling
loads for computerizing energy calculations. Algorithms for building
heat transfer subroutines, ASHRAE, New York, 1975, pp. 7678.
[38] K.K. Tien, E.M. Sparrow, Local heat transfer and uid ow
characteristics for airow oblique or normal to a square plate, Int.
J. Heat Mass Transfer 22 (1979) 349360.
[39] B. Jennings, Environmental Engineering, International Textbook
Company, 1970.
[40] N.S. Sturrock, Localized boundary layer heat transfer from external
building surfaces, Ph.D. Thesis, University of Liverpool, 1971.
[41] K. Kimura, Scientic Basis of Air Conditioning, Applied Science
Publishers Ltd., London, England, 1977.
[42] K. Nicol, The energy balance of an exterior window surface, Inuvik,
N.W.T., Canada, Build. Environ. 12 (1977) 215219.
[43] CIBS, Chartered Institute of Building Services Guide Book A, Section
A3, CIBS, London, 1979.
[44] F.L. Test, R.C. Lessmann, A. Johary, Heat transfer during wind ow
over rectangular bodies in the natural environment, Trans. ASME J.
Heat Transfer 103 (1981) 262267.

[45] M. Yazdanian, J.H. Klems, Measurement of the exterior convective


lm coecient for windows in low-rise buildings, ASHRAE Trans.
100 (1) (1994) 10871096.
[46] S. Sharples, P.S. Charlesworth, Full-scale measurements of windinduced convective heat transfer from a roof-mounted at plate solar
collector, Solar Energy 62 (1998) 6978.
[47] H.H. Sogin, A summary of experiments on local heat transfer from
the rear of blu obstacles to a low speed airstream, Trans. ASME J.
Heat Transfer (1964) 200202.
[48] J.W. Mitchell, Base heat transfer in two-dimensional subsonic fully
separated ows, Trans. ASME J. Heat Transfer (1971)
342348.
[49] N. Onur, J.C. Hewitt Jr., A study of wind eects on collector
performance, in: Proceedings of the ASME Solar Energy Division,
Century 2 Solar Energy Conference, San Francisco, CA, August 19
21, 1980.
[50] A.H. Taki, D.L. Loveday, External convection coecients for framed
rectangular elements on building facades, Energ. Build. 24 (1996)
147154.
[51] S. Shakerin, Wind-related heat transfer coecient for at-plate
solar collectors, Trans. ASME J. Solar Energy Eng. 109 (1987)
108110.
[52] N. Onur, Forced convection heat transfer from a at-plate model
collector on roof of a model house, Warme und Stoubertragung 28
(1993) 141145.
[53] http://www.eere.energy.gov/buildings/energyplus/.
[54] http://sel.me.wisc.edu/trnsys/.
[55] http://www.esru.strath.ac.uk/Programs/ESP-r.htm.
[56] S.J. Rees, D. Xiao, J.D. Spitler, An analytical verication test suite
for building fabric models in whole building energy simulation
programs, ASHRAE Trans. 108 (1) (2002) 3042.
[57] F.C. McQuiton, J.D. Parker, J.D. Spitler, Heating, Ventilating, and
Air Conditioning, Analysis and Design, fth ed., John Wiley & Sons
Inc., New York, 2000, p. 226.
[58] D. Bradley, TRNSYS engineer, private communication.
[59] J.S. Coventry, in: Proceedings of the 40th ANZSES Solar Energy
Conference, 2002.
[60] ASHRAE, Handbook of Fundamentals, SI Edition, Atlanta, GA,
USA, 1997.
[61] http://www.esru.strath.ac.uk/Programs/ESP-r_CodeDoc/esrubld/
convect2.F.html.
[62] ESP-r Ver. 9, Data Model Summary, 5.6, ESRU, University of
Strathclyde, 2001.
[63] L. Zhang et al., Research on system identication of wall surface heat
transfer processes, Exp. Heat Transfer 15 (2002) 3147.
[64] C. Cristofari et al., Thermal modeling of a photovoltaic module, in:
Proceedings of the 6th IASTED International Conference Modeling,
Simulation, and Optimization, September 1113, 2006, Gaborone
Botswana, pp. 273278.
[65] T. Schott, Operation temperatures of PV modules, in: W. Palz, F.C.
Treble (Eds.), Proceedings of the 6th E.C. PV Solar Energy
Conference, London, 1519 April 1985, D. Reidel Publ. Co., 1985,
pp. 392396.
[66] C. Wray, Duct thermal performance models for large commercial
buildings, Lawrence Berkeley National Laboratory Report, LBNL53410, 2003.
[67] P. Nolay, Developpement dune methode generale danalyse des
systemes photovoltaques. MS Thesis, Ecole des Mines, SophiaAntipolis, France, 1987.
[68] K. Furushima, Y. Nawata, M. Sadatomi, Prediction of photovoltaic
(PV) power output considering weather eects, in: Proceedings of the
SOLAR 2006 Renewable Energy, Key to Climate Recovery, July 7
13 2006, Denver, Colorado, USA.
[69] L. Wen, Investigation of the eect of wind cooling on photovoltaic
arrays, Report DOE/JPL-1012-69, 1982.

Anda mungkin juga menyukai