Sie sind auf Seite 1von 37

2.

SURFACE TRANSFER COEFFICIENTS


2.1 INTRODUCTION

In building simulation, transport phenomena, as air flow, heat transfer or mass transfer,
are modelled inside buildings, between bodies (walls) and air, at the outside of buildings,
… Boundary conditions are represented by defining a transfer of a thermodynamic
property (flux) between these walls and the internal or external air flow, or by defining a
fixed state at the wall. In order to model the interaction between the wall (state) and the
fluid (state) a transfer coefficient is often used, known as friction coefficient, heat transfer
coefficient or mass transfer coefficient.
Different authors have examined the sensitivity of thermal predictions from energy
simulation programs to the modelling of internal convection (e.g. Spitler et al. (1991),
Clarke (1991), Fisher and Pederson (1997)). Their work has demonstrated that
predictions of energy demand and consumption can be strongly influenced by the choice
of (made by program developer or user) heat transfer calculation method. Differences of
20-40% in energy predictions were noted by some of these authors.
More importantly, the predicted benefits from design measures were, in some cases,
found to be sensitive to the approach used to model internal surface convection. As a
result, the choice of heat transfer calculation method could affect the design decisions
drawn from a simulation-based analysis. (Beausoleil-Morrison (1999)).

The transfer coefficient is in fact a modelling assumption in itself. The concept of transfer
coefficients is developed in the boundary layer theory, first derived by Ludwig Prandtl in
1904. Prandtl discovered that for most applications the influence of viscosity is confined
to an extremely thin region very close to the body and that the remainder of the flow
could, to a good approximation, be treated as inviscid. The pressure in the boundary
layer and in the main flow is assumed to be the same.
This clearly shows that transfer coefficients are by nature an, though often good,
approximation. They should be used within the constraints of the approximation. They
are only applicable for the correct boundary conditions. As simulations advance to
include more details, the improper use of transfer coefficients often leads to non-physical
results.
In this chapter the basic concepts of boundary layer theory are introduced and the main
parameters describing friction, heat and mass transfer are addressed. For further review
reference is made to Kays & Crawford (1993) and Welty et al (2001).
Application to buildings is discussed through papers published during ANNEX 41 and
recent publications in literature.

2.2 BOUNDARY LAYER THEORY IN A NUTSHELL

2.2.1 Convection - flux laws

Among the many tasks facing the engineer is the calculation of energy-transfer and
mass-transfer rates at the interface between phases in a fluid system. Most often we are
concerned with transfer at a solid-fluid interface where the fluid may be visualised as
moving relative to a stationary solid surface.
If the fluid is at rest in the entire domain, the problem becomes one of simple heat
conduction where there are temperature gradients normal to the interface or simple
mass diffusion where there are mass concentration gradients normal to the surface.
However, if there is fluid motion, energy and mass are transported both by potential
gradients (as in simple conduction) and by movement of the fluid itself. This complex
transport process is usually referred to as convection.

Flow
tws v
t
q&

cws c

m&

Figure 1 : Heat and mass transfer from a surface in contact with a fluid

In simple convective heat transfer along a wall it is often convenient to define a


convection heat-transfer conductance or heat transfer coefficient as (Figure 1):

q& = h(t ws − t ∞ )

The driving force for heat transfer ( q& ) is the temperature difference between the wall
surface (tws) and the free fluid stream (t∞). This equation is also known as Newton’s Law
of Cooling.
The conductance h is in essence a fluid mechanic property of the system and t,
temperature, a thermodynamic property. There are numerous non-linear applications
were h is itself a function of the temperature difference. It is important to note that in that
case the equation remains valid as a definition of h, although it may well reduce the
usefulness of the conductance concept.

In mass transfer it is convenient to define a convective mass-transfer conductance such


that the total mass flux at the surface ( m
& )is the product of the conductance g and the
driving force, being the difference in concentration at the wall (cws) and in the free fluid
stream (c∞).
m& = hm (cws − c∞ )
The conductance hm is essentially a fluid mechanic property of the system, whereas the
concentration is a thermodynamic property.

This second equation has the same form as the first one resulting to the rise of the heat-
and mass- transfer analogies.

The form of these equations is in fact a special case of the general form of a convection
coefficient as given by :

FLUX = COEFFICIENT x Driving Force

2.2.2 Hydraulic, thermal and concentration boundary layer

Figure 2 : Heat and mass transfer from a surface in contact with a fluid

In 1904, Ludwig Prandtl stated : “At high Reynolds number the effect of fluid friction is
limited to a thin layer near the boundary of the body” , hence the term THE BOUNDARY
LAYER came into engineering practice.

Figure 2 shows the boundary layer developing over a flat plate under forced convection,
meaning there is an external velocity v∞ which is causing the flow over the plate. This
velocity can be created by a fan, wind, …. The thickness of the boundary layer (δ) is
arbitrarily taken as the distance away from the surface where the velocity reaches 99%
of the free stream velocity. Figure 2 illustrates how the thickness of the boundary layer
increases with distance x from the leading edge. At relatively small values of x flow
within the boundary layer is laminar. At larger values of x the transition region is shown
where fluctuations between laminar and turbulent flow occur within the boundary layer.
Finally above a certain value of x the boundary layer will always be turbulent. In the
turbulent boundary layer a small laminar sublayer exists were there are steep velocity
gradients.
The criterion for the type of boundary layer is the magnitude of the Reynolds number:

v∞ x
Rex =
ν

In general is the Reynolds number is lower than a certain value, depending on the
geometry flow, is laminar. Above a second value for the Reynolds number the flow is
fully turbulent. In between transitional flow occurs. In general a Reynolds number is
defined as

vL
Re =
ν
with v a characteristic velocity in the flow and L a characteristic length.

By solving the Navier-Stokes equations for a two dimensional flow for this geometry, as
discussed in Kays WM, Crawford ME, (1993), the hydraulic boundary layer thickness as
function of the position along the plate can be found. The hydrodynamic or momentum
boundary layer may be defined as the region in which the fluid velocity changes from it’s
99% free stream value to zero at the body surface. This is not a precise definition of the
boundary layer thickness. It only means that the boundary layer thickness is the distance
from the wall in which most of the velocity change takes place.
Out of this analysis follows the drag coefficient also known as the friction coefficient (cf):
τ
cf =
ρv / 2
2

with ρ the fluid density and τ the fluid friction or shear stress.

When there is heat or mass transfer between the fluid and the surface, it is also found
that in most practical applications the major temperature and concentration changes
occur in the region very close to the surface. This gives rise to the concept of the thermal
boundary layer and the concentration boundary layer, and again the relative thinness of
these boundary layers permits the introduction of boundary-layer approximations similar
to those introduced for momentum. Solving the Navier-Stokes equations for the energy
or concentration transport equations results in a thermal boundary layer thickness and a
concentration boundary layer thickness as function of the coordinate x.
μc p
In the solution of the diffential equations the Prandtl number Pr = appears, relating
k
the viscous boundary layer to the thermal boundary layer. For mass transfer this is
μ
expressed by the Schmidt number Sc = relating the viscous boundary layer to the
ρD AB
concentration boundary layer.
If the ratio is taken of the Prandtl number to the Schmidt number the Lewis number is
Pr
found, relating mass to thermal diffusion Le = . As this number relates the thermal to
Sc
the mass transfer boundary layer it will determine the analogy between heat and mass
transfer.
In the 19th century Reynolds was the first to report on the analogous behaviour of heat
and momentum transfer (Welty et al. 2001). He presented results on frictional resistance
to fluid flow in conduits which made the quantitative analogy between the two transport
phenomena possible. Out of these observations the Reynolds analogy was stated. The
Reynolds analogy relates the heat transfer coefficient (h) to the skin friction coefficient
using the free stream velocity and the free stream density and heat capacity (cp):
h cf
St = =
ρv∞ c p 2
This relation can be deduced out of the boundary layer equations for laminar forced flow
across a solid boundary under the conditions that the Prandtl number (Pr) is equal to
one and no form drag is present.
The Reynolds analogy can also be applied to mass transfer in case the Schmidt number
(Sc) is equal to one:

hm cf
=
v ∞c p 2
In case both Pr and Sc numbers are equal to one, and hence the Lewis number (Le) is
one. Comparing both equations, a relation between the mass transfer coefficient and the
heat transfer coefficient is found, hence the analogy between heat and mass transfer
was founded :

h
= hm
ρv∞ c p
In general the convection heat transfer coefficient is made dimensionless through the
definition of a Nusselt-number and the mass transfer convection coefficient through the
definition of the Sherwood-number

hL
Nu =
k

hm L
Sh =
D AB

For forced convection the heat and mass transfer coefficients can be expressed as the
Nusselt number as function of the Reynolds and Prandtl number :

Nu = F (Re , Pr )
Sh = F (Re ,Sc )

For natural convection the flow is driven by buoyancy as a result of density differences in
the air volume. The dimensionless number characterising this flow type is the Grashof
number given by

gΔρL3
Gr =
ρν 2
For natural convection the Grashof number takes over from the Reynoldsnumber to
determine the convection coefficients :

Nu = F (Gr , Pr )
Sh = F (Gr , Sc )

For the convective heat transfer coefficient a lot of data is available. For several, relative
simple geometries and different flow conditions (laminar, transitional, turbulent, forced
and buoyancy driven convection) an analytical solution of the Navier-Stokes equations
applied to a boundary layer exists. (See eg Kays WM, Crawford ME, 1993)
For more complex geometries correlations have be determined by curve fitting
dimensionless numbers to large data sets.
As there are many different correlations available care has to be taken in selecting the
suitable correlation. For analytical derived correlations the validity of assumptions and
simplifications should be checked. For experimentally derived correlations the range and
accuracy of the data set should be taken into consideration.

For the mass transfer coefficient boundary layer analysis leads again to analytical
solutions. Due to the fact that the differential equations for heat and mass transfer
resulting from boundary layer analysis are analogues, the solutions obtained for heat
transfer can be transformed into mass transfer solutions, by using the correct
dimensionless number cited earlier (Welty et al (2001)).
Furthermore it is very difficult to determine the convective mass transfer coefficient
experimentally. Therefore this analogy is applied in a lot of cases for calculating the
convective mass transfer coefficient, starting from the thermal measurements that where
done. Validity of the thus obtained mass transfer coefficients is by consequence even
more limited and great care should again be taken in selecting the proper correlation for
the studied geometry. (See eg Kays WM, Crawford ME, 1993)

For flow around buildings very little information was found about mass transfer
determination, both experimentally or numerically. For flows inside buildings, most
research is focussing on flows over building materials or porous materials. Wadso,L ,
1993 gives a very broad literature review.

During the progress of the Annex 41 new experiments were proposed to determine the
mass transfer coefficient. Often these experiments were found to have limited validity.
Secondly numerical methods, based on CFD, were used to determine mass transfer
from a fluid to a porous material. Finally the heat and mass transfer analogy was looked
into.
2.2.3 Overview of useful dimensionless numbers and nomenclature

Name Symbol Definition Meaning


Reynolds number Re vL inertial forces compared to
Re = viscous forces
ν
Grashof number Gr gΔρL3 buoyancy forces compared to
Gr = viscous forces
ρν 2
Nusselt number Nu hL dimensionless heat transfer
Nu = coefficient
k
Sherwood number Sh h L dimensionless mass transfer
Sh = m coefficient
D AB
Prandtle number Pr μc p viscosity to thermal diffusion
Pr =
k
Schmidt number Sc μ viscosity to mass diffusion
Sc =
ρD AB
Stanton Number St h Nu ratio of heat transferred into a
St = = fluid to the thermal capacity of
ρv∞ c p Re Pr fluid
Lewis number Le Pr thermal diffusion to mass
Le = diffusion
Sc
Archimedes number Ar gρ l (ρ l − ρ g )L3 the motion of fluids due to density
Ar = differences
μ2
Rayleigh number Ra Ra = Gr Pr
Richardson number Ri Gr Measure for comparing forced
Ri = convection to buoyancy driven
Re 2 convection (mixed convection)
Units
Heat capacity cp J/kgK
Diffusion coefficient DAB m²/s
Gravitational g 9.81 m/s²
constant
Heat transfer h W/m²K
convection
coefficient
Mass transfer hm m/s
convection
coefficient
Characteristic Length L M
Flow velocity v m/s
Free stream velocity v∞ m/s
Boundary layer δ m
thickness
Thermal conductivity λ W/mK
Dynamic viscosity μ Pa s
Kinematic viscosity ν m²/s
Density ρ kg/m³
Table 1 : Dimensionless numbers and nomenclature
2.3 HEAT TRANSFER

2.3.1 Flow over and around buildings (experimental data)


Air flows around buildings are mainly of a forced nature as they are caused by wind.
Exterior convective heat and mass transfer coefficients at building surfaces are to a large
extent determined by the local wind speed. Usually, empirical formulae are used to relate
the reference wind speed at a meteorological station to the local wind speed near the
building surface and to relate the local wind speed to surface transfer coefficients. These
formulae however are based on a limited number of measurements.

Practical correlations given by Jürges (1924) give a relation between free stream wind
( V∞ ) speed and the thermal convection coefficient :

h = 4.0V∞ + 5.6 ; V∞ < 5 m / s


h = 7.1V∞0.78 + 5.6 ; V∞ > 5 m / s

Charples (1984) presented the following algorithm :

h = 1.7Vloc + 5.1

with Vloc the local wind speed measured at 1 m distance from the surface. It is expressed
as a simple function of the reference wind speed U10 :

Vloc = 1.8U 10 + 0.2 ; windward


Vloc = 0.4U 10 + 1.7 ; leeward

Also ASHRAE (ASHRAE 1975) proposed practical correlations of a similar nature :

h = 1.886U loc
0.605

windward U loc = 0.25U 10 if U 10 > 2m / s


= 0.5 if U 10 < 2m / s
leeward U loc = 0.05U 10 + 0.3

These type of correlations are proven to be very sensitive to errors. First of all the
definition of the free stream wind speed around a building is not clear and secondly the
relation between wind speed and the actual heat transfer coefficient is not clearly stated.
This is discussed in Annex paper A41-T3-B-05-5.

2.3.2 Flow over and around buildings (CFD data)

Little is known about the actual value and the variability of local wind speed and surface
transfer coefficients across facades of different building geometries. In Annex paper
A41-T3-B-05-5 a validated Computational Fluid Dynamics (CFD) model (FLUENT) is
used to calculate the local wind speed near the exterior surface of a cubic building model
as a function of wind speed, wind direction and the position on the facade.
It is shown that the variation of the local wind speed across the facade is very
pronounced and that using the available empirical formulae can yield large errors in
HAM calculations.

In Annex paper A41-T3-Br-07-2 (and Emmel M, Abadie M, Mendes N (2007)) similar


conclusions were drawn using in essence the same approach. Using CFD calculations
with CFX, correlations for the heat transfer coefficient were determined for the BESTEST
reference case Judkoff R.D., Neymark J.S. (1995). De correlations presented in the
previous paragraph were compared to the CFD calculations and both over and under
predictions (to about a factor 4) of these correlations were found in relation to the CFD
solutions. The paper ends with a list of new correlations determined by doing several
calculations with CFD on the BESTEST geometry. These are copied here. More
information about validity and boundary conditions can be found in the paper.

Table 2 : Data according to A41-T3-Br-07-2

2.3.3 Flow inside buildings (experimental data)


Air flows inside buildings occur due to two main reasons. Firstly there are air streams
caused by ventilation systems (jets) or pressure differences between adjacent rooms
(draught). These are thus of forced nature as the flow is not driven by the temperature or
density fields it creates. Secondly temperature and concentration (vapour) differences
inside a room cause density differences and thus buoyancy.
Inside buildings both forced and natural convection will occur. Sometimes they will even
operate at the same place and time. This is what is called mixed convection.

2.3.3.1 Forced convection inside buildings


Spitler et al. (1991b) designed a full-scale experimental facility, with internal dimensions
of 4.57 x 2.74 x 2.74 m and a fan system delivered air to one of the two room inlets over
a range of 5 to 100 air changes per hour (ACH). The walls, floor and ceiling were
covered by heated panels, each with an independent electrical resistance heater.
Figure 3 Experimental facility of Spitler et al. 1991a

Spitler et al. correlated the convective heat transfer coefficients to the jet momentum
number J:
m& U 0
h = C1 + C 2 ⋅ J 0.5 with J = (U0 jet inlet velocity, Vroom room volume)
ρgVroom
The correlations from Spitler et al. (1991b) are listed in the Table below.

⎛ gβΔTL ⎞
Table 3.Heat transfer coefficient correlations of Spitler et al. (1991b) ⎜⎜ Ar = ⎟
⎝ U 02 ⎟⎠
Surface Inlet h Limits
Ceiling Ceiling 11.4 + 209.7 J0.5 0.001<J<0.03
Vertical walls Ceiling 4.2 + 81.3 J0.5 0.001<J<0.03
Floor Ceiling 3.5 + 46.8 J0.5 0.001<J<0.03
Ceiling Side wall 10.6 + 59.4 J0.5 0.002<J<0.011 and Ar<0.3
Vertical walls Side wall 1.6 + 92.7 J0.5 0.002<J<0.011
Floor Side wall 3.2 + 44.0 J0.5 0.002<J<0.011 and Ar<0.3

Fisher extended Spitler’s work by investigating buoyant, wall and free jets over a range
of room inlet conditions using the same enclosure. The room was also isothermal in
most of the experiments. However, a single wall was chilled in one group of ceiling-
diffuser experiments to examine the combined influence of buoyancy and forced effects.
The experiments only examined room cooling, the incoming air stream always being
colder than the air within the room and the room surfaces.
Correlations were developed for three classes of flows: (1) ceiling jets in isothermal
rooms, (2) free horizontal jets in isothermal rooms and (3) ceiling jets in non-isothermal
rooms.
For the first class, Fisher (1995) found surface convection to be independent of the inlet
velocity of the ceiling jet, but rather to depend upon the jet’s volumetric flow rate. He also
found the buoyancy forces of the cold jet to be negligible relative to the viscous Coanda
effect, adhering the jet to the ceiling and walls. The form of the correlations, expressed in
dimensionless parameters, can be described as follows: (applicable 3< ACH< 100)
hVroom
1/ 3
Nu = C1 + C 2 ⋅ Re C3
e where Nu =
λ
ρV&diffuser
and Re e = ( V&diffuser jet volumetric flow rate m³/s)
μV 1/ 3
room

With the free horizontal jets in isothermal rooms the buoyancy forces of the cold jet also
had a negligible impact on convection from the walls and floor. Therefore the same type
of equation was also used to correlate these data. Convection from the ceiling, though,
was affected by buoyancy.
Consequently, an alternate equation to correlate the ceiling data: (applicable for 3<
ACH< 12)

Re Ce3
Nu = C1 + C 2 ⋅
Are

Table 4 : Convective heat transfer correlations of Fisher (1995)


Surface type Configuration h
Walls − 0.199 + 0.190 ⋅ ( ACH )
0.8

0.159 + 0.116 ⋅ ( ACH )


Floor Forced convection with ceiling 0.8
diffuser in isothermal rooms
Ceiling − 0.166 + 0.484 ⋅ ( ACH )
0.8

Walls − 0.110 + 0.132 ⋅ ( ACH )


0.8

0.704 + 0.168 ⋅ ( ACH )


Floor Forced convection with wall 0.8
diffuser (free jet)
Ceiling 0.064 + 0.00444 ⋅ ( ACH )
0.8

Fisher and Pedersen (1997) correlated the same data as Fisher (1995) using a different
functional form. The correlations are applicable for forced convection and work in similar
conditions as the correlation developed by Fisher (1995). The main difference is the
position of the diffuser. These correlations are applicable in rooms with ceiling diffusers
where the diffuser jet is attached to the ceiling surface. As with previous correlations,
supply air temperature is used as the reference temperature. The correlations are given
in Table 5.

Table 5 :Convective heat transfer correlations of Fisher and Pedersen (1997)


Surface type Configuration h
Walls 0.19 ⋅ ( ACH )
0.8

0.13 ⋅ ( ACH )
Floor Forced convection with ceiling 0.8
diffuser in isothermal rooms
Ceiling 0.49 ⋅ ( ACH )
0.8

2.3.3.2 Natural convection or buoyancy driven flow inside buildings


Alamdari and Hammond (1983) are one of the first to develop correlations dedicated to
building applications. Correlations that cover laminar, transitional and turbulent flow
regimes for the following three configurations are given: (1) vertical surfaces, (2) stably-
stratified horizontal surfaces (e.g. warm air above a cool floor) and (3) buoyant flow from
horizontal surfaces (e.g. cool air above a warm floor). The correlations cover the full
range of temperature and dimensions that appear in buildings. But they are not
applicable for cases where buoyancy is created by thermal devices (e.g. radiator
baseboard heater or fan coil) and mechanically driven jets as experienced in
mechanically ventilated buildings. They are only valid for purely buoyant flow, in cases
where buoyancy is caused by the temperature difference between a surface and the
surrounding room air (ΔT). Alamdari and Hammond (1983) did not perform new
experiments, the correlations are based on collected experimental data reported in the
literature. All data is derived from experiments conducted with free standing surfaces
(surfaces not part of the room), which limits the applicability of the correlations.

Table 6 : Alamdari and Hammond convection correlations (Beausoleil-Morrison, 2000)


(H is room height en Dh the room hydraulic diameter)
Surface type Ventilation regime h
Wall ⎧ 1/ 4 6 ⎫
1/ 6

⎪⎡ ⎤
⎛ ΔT ⎞
⎨⎢1.5 ⋅ ⎜
H
⎟ [
⎥ + 1.23ΔT
1/ 3
] ⎪⎬
6


⎪⎩⎣ ⎝ ⎠ ⎥⎦ ⎪⎭
Floor (Tsurface>Tair) 1/ 6
⎧⎡ 1/ 4 6
⎤ ⎫
⎛ ⎞
Δ
[ ]
Ceiling (Tsurface < Natural convection ⎪ T 6⎪
Tair) (system is off) ⎨ ⎢1.4 ⋅ ⎜ ⎟
⎜ D ⎟ ⎥ + 1.63ΔT
1/ 3

⎪⎩⎢⎣ ⎝ h ⎠ ⎥⎦ ⎪⎭
Floor (Tsurface < Tair) ⎛ ΔT ⎞
1/ 5

Ceiling (Tsurface>Tair) 0.6 ⋅ ⎜⎜ ⎟⎟


⎝ Dh ⎠

Khalifa and Marshall (1990) performed experiments in a room sized test cell to produce
correlations specific to internal convection within buildings. Convection correlations are
developed based on measurements in an experimental chamber with room sizes: 2.95 x
2.35 x 2.08 m (l x w x h). The correlations for vertical surfaces are defined for surfaces in
the vicinity of a terminal device and for other surfaces. To assess a number of common
convection regimes, the test cell’s configuration was varied. Different heating systems
(e.g. radiator, in-floor heating, convective heating) were analyzed, as was the placement
of the heating device (e.g. underneath a window or facing a window).

Figure 4 : Experimental test room of Khalifa and Marshall (1990)

Khalifa (1989) used the average room air temperature as the reference temperature to
calculate the convective heat transfer coefficient. But Khalifa and Marshall (1990)
measured the air temperature outside the thermal boundary layer at a distance of 60 mm
from the interior surface of the wall, which is used as the reference temperature.
Khalifa generated a total of 36 correlations. Data from similar correlations were
combined together in order to obtain new and more general correlations which can be
applied in more than one configuration (Khalifa and Marshall 1990). By combining these
similar results the data were collapsed into a series of 10 equations (Tables 7 and 8).

Table 7 Khalifa convection correlations (Beausoleil-Morrison 2000)


Surface type Ventilation regime h
Wall Rooms heated by radiator
In the vicinity of the Radiator not located under window 1.98 ⋅ ΔT 0.32
terminal device Only surfaces adjacent to radiator
Wall Rooms heated by radiator
Radiator located under window 2.3 ⋅ ΔT 0.24
Rooms with heated walls
Not applicable for heated walls
Wall Rooms heated by circulating fan
heater 2.92 ⋅ ΔT 0.25
Only for surfaces opposite to fan
Wall Rooms heated by circulating fan
heater
For surfaces not opposite to fan 2.07 ⋅ ΔT 0.23
Rooms with heated floor
Rooms heated by radiator
Radiator not located under window
For surfaces not next to radiator
Window Rooms heated by radiator
Radiator located under window 8.07 ⋅ ΔT 0.11
Window Rooms heated by radiator
Radiator not located under window 7.61 ⋅ ΔT 0.06
Ceiling Rooms heated by radiator
Radiator located under window
3.1 ⋅ ΔT 0.17
Rooms with heated walls
Ceiling Rooms heated by circulating fan
heater
Rooms with heated floors 2.72 ⋅ ΔT 0.13
Rooms heated by radiator
Radiator not located under window

Table 8 Khalifa and Marshall (1990) convection correlations


Surface type Ventilation regime h
Wall
Large isolated vertical surface 2.03 ⋅ ΔT 0.14
Floor
Large heated surface facing upward 2.27 ⋅ ΔT 0.24
Calay et al. (1998) performed an experimental study of buoyancy-driven convection in
rectangular enclosures. The enclosure was one-quarter scale model of a typical room. It
was based on ‘hot box’ arrangement, in which two opposing walls are heated and cooled
while others are insulated and act as adiabatic walls. Four sets of experiments were
performed to simulate the following convective heat-flow configurations: (1) enclosure
heated from side, (2) large vertical walls as hot and cold plates, (3) small vertical walls
as hot and cold plates, (4) enclosure heated from below, (5) stably stratified convection
(enclosure heated from ceiling).
The convective heat transfer correlations are given in terms of dimensionless
parameters: Nusselt, Prandtl and Grashof number. The correlations recommended by
ASHREA (1985) and CIBSE (1986) and other correlations derived from tests with full
size enclosures and similar configurations are used for comparing the experimental
results (Table 9)

Table 9 Equations employed for comparison (Calay et al. 1998)


Equation Correlation, Nu Gr range Flow
condition
108<Gr<101
Configuration: stably 0
stratified, Tw=cte
0.236Gr 1/4
10 <Gr<101
8
Laminar
CIBSE (1986) 0
0.218Gr1/4 Laminar
ASHRAE (1985)
0.56Gr1/5 108<Gr<101 Laminar
Alamdari and Hammond 0
0.065Gr0.255 Not specified
(1983)
Not
Min et al. (1956)
specified

Awbi and Hatton (1999) conducted experiments in two experimental chambers, with
different size in order to assess scale effects (Figure 5). The first chamber had a typical
room size of 2.78 x 2.78 x2.3 m. The second was considerably smaller 1 x 2.78 x 2.3 m.
This chamber was kept at a low temperature, so that the wall that connected the two
chambers acted as a heat sink. The main chamber was conditioned by electrically
heated plates affixed to the surfaces. A single surface (wall, floor or ceiling) was heated
in each experiment. All the walls were aluminium plated and long-wave radiation was
taken into account.

Figure 5 Environmental chamber of Awbi and Hatton (1999)


Natural convection from heated room surfaces was characterized by a correlation of the
mean convection heat transfer coefficient for whole wall heated surfaces.

Table 10 Awbi and Hatton (1999) natural convection correlations


Surface type Ventilation Nu
regime
Walls
0.289(Gr )
0.293
( 9 x 108 < 6 x 1010)
Floors 0.269(Gr )
0.308

( 9 x 108 < 7x 1010)


Buoyant with
Ceilings heated surface 1.78(Gr )
0.133
( 9 x 108 < 1 x 1011)
Partly heated
3.517(Gr )
0.16
ceilings

Khalifa (2001) gives an extensive review of studies about natural convective heat
transfer coefficients on surfaces in two- and three-dimensional enclosures with primary
focus on those with a direct application to heat transfer in buildings. Figures 6 to 8 give a
comparison of the different correlations mentioned in Khalifa (2001).

Figure 6 : Convective heat transfer coefficient correlations for vertical surfaces (Khalifa
2001)
Figure 7 Convective heat transfer coefficient correlations for heated plate facing upward
(heated floor/ cold ceiling) (Khalifa 2001)

Figure 8 Convective heat transfer coefficient correlations for heated plate facing
downward (heated roof/cold floor) (Khalifa 2001)

2.3.3.3 Mixed convection inside buildings


Beausoleil-Morrison (2000) developed a suitable method for solving mixed flow. He
created his correlations by combining the correlations for natural convection (Alamdari
and Hammond 1983) and for forced convection where the air is supplied by a ceiling
diffuser (Fisher 1995). In some cases the mechanical and buoyant forces will assist (act
in same direction, Figure 9) while in others they will oppose (act in opposite directions) or
act transversely (act in perpendicular directions). It is difficult (usually impossible) to
predetermine whether a configuration will be dominated by buoyant forces or mechanical
forces. Beausoleil-Morrison solves this problem by selecting and combining the
appropriate correlations for forced and natural convection.
Figure 9 Assisting mechanical and buoyant forces

Table 11 Convective heat transfer coefficient correlations of Beausoleil-Morrison (2000)


for mixed flow
Surface type h
1
Wall ⎛⎧ 3× 1 ⎞
⎜ ⎡ ⎛ Δθ ⎞1 / 4 ⎤
6
⎫ 6

[ ]
3
⎧⎪ ⎡ θ − θ ⎫
Assistin

⎜ ⎨ ⎢1 .5⎜ ⎟ ⎥ + 1 .23 Δ θ 1 / 3
⎜ ⎪ ⎢ ⎝ H ⎟⎠ ⎥

6⎪
⎬ + ⎨⎢ s d

[
⎥ − 0 .199 + 0 .190 ( ACH )0.8 ]⎪⎬ ⎟

⎪ ⎪⎩ ⎢⎣ Δ θ ⎥⎦ ⎪⎭
g forces ⎜ ⎩⎣ ⎦ ⎭ ⎟
⎝ ⎠

⎡⎛ 3× 1
⎢ ⎜ ⎧⎪ ⎡ ⎛ ⎫
6 6

[ ]
3
⎧⎪ ⎡ θ − θ ⎫⎪

[ ]
1/ 4
Δθ ⎞ ⎪
⎥ − 0 . 199 + 0 . 190 ( ACH )
6
⎢ ⎜ ⎨ ⎢1 . 5 ⎜ ⎟ ⎥ + 1 . 23 Δ θ − ⎨⎢ s d
1/ 3 0 .8
⎬ ⎬
⎢ ⎜ ⎪ ⎢⎣ ⎜⎝ H ⎟⎠ ⎥⎦ ⎪ ⎪⎩ ⎢⎣ Δ θ ⎥⎦ ⎪⎭
⎢ ⎜⎝ ⎩ ⎭
⎢ 1/ 6
⎢ ⎧⎡ 1/ 4 6 ⎫
Opposin max ⎢ 80 % of


⎨ ⎢1 . 5
⎛ Δθ ⎞ ⎤



⎟ ⎥ + [
1 . 23 Δ θ
1/ 3 ⎪
⎬ ]
⎢ ⎪ ⎢⎣ ⎝ H ⎠ ⎥⎦ ⎪
g forces ⎩ ⎭

⎧⎪ ⎡ θ s − θ d ⎤ ⎫

⎢ 80 % of ⎨⎢ [
⎥ ⋅ − 0 . 199 + 0 . 190 ( ACH )0 .8 ]⎪⎬
⎢ ⎪⎩ ⎢⎣ Δ θ ⎥⎦ ⎪⎭

⎣⎢

Floor ⎛⎧ 3× 1 ⎞
1/ 3

⎜ ⎡ ⎛ Δθ ⎞1 / 4 ⎤
6
⎫ 6

[ ]
3
⎧⎪ ⎡ θ − θ ⎫

⎜ ⎨ ⎢1 .4 ⎜ ⎟ ⎥ + 1 .63 Δ θ 1 / 3
⎜ ⎪ ⎢ ⎝ D h ⎟⎠ ⎥

6⎪
⎬ + ⎨⎢ s d

[
⎥ 0 .159 + 0 .116 ( ACH )0.8 ]⎪⎬ ⎟

Buoyant ⎪ ⎪⎩ ⎢⎣ Δ θ ⎥⎦ ⎪⎭
⎜ ⎩⎣ ⎦ ⎭ ⎟
⎝ ⎠

1/ 3
⎛⎡ 1/ 5 3 ⎞
⎛ Δθ ⎞ ⎤
3
⎧⎡ θ −θ d ⎫⎪
Stably ⎜
⎜ ⎢ 0 .6⎜⎜ ⎟ ⎥ + ⎪⎨ ⎢ s


[
⎥ ⋅ 0 .159 + 0 .116 ( ACH ) 0 .8 ]⎬


stratified ⎜ ⎢⎣ ⎝ D h ⎠ ⎥⎦ ⎪⎩ ⎢⎣ Δ θ ⎦⎥ ⎪⎭ ⎟
⎝ ⎠
1
Ceilin ⎛⎧ 3× 1 ⎞
⎜ ⎡ ⎛ Δθ ⎞ 1 / 4 ⎤
6
⎫ 6

[ ]
3
⎧⎪ ⎡ θ − θ ⎤ ⎫
g
Buoyant

⎜ ⎨ ⎢1 .4⎜ ⎟ ⎥ + 1 .63 Δ θ 1 / 3
⎜ ⎪ ⎢ ⎜⎝ D h ⎟⎠ ⎥
6⎪
⎬ [
+ ⎨ ⎢ s d ⎥ − 0 .166 + 0 .484 ( ACH )0.8 ]⎪⎬ ⎟

⎪ ⎪⎩ ⎣⎢ Δ θ ⎥⎦ ⎪⎭
⎜ ⎩⎣ ⎦ ⎭ ⎟
⎝ ⎠
1/ 3
⎛⎡ 1/ 5 3 ⎞
⎛ Δθ ⎞ ⎤
3
⎧⎪ ⎡ θ s − θ d ⎤ ⎫⎪
Stably ⎜
⎜ ⎢ 0 . 6 ⎜⎜ 2 ⎟⎟ ⎥ + ⎨ ⎢ [
⎥ ⋅ − 0 . 166 + 0 . 484 ( ACJH )0 .8 ]⎬


stratified ⎜ ⎢⎣ ⎝ D h ⎠ ⎥⎦ ⎪⎩ ⎣⎢ Δ θ ⎦⎥ ⎪⎭ ⎟
⎝ ⎠

The experiments of Awbi and Hatton (2000) were carried out in the same enclosure as
the natural convection experiments. They only placed an air handling unit onto the
ceiling of the small (cold) compartment to cool the dividing wall that separates the two
compartments. The fan and heating plates were positioned on a wall, the floor and the
ceiling to investigate the effect of a 3D wall jet on the surface convective heat transfer
coefficient (Figure 10). The flow regime was a combination of natural convection,
caused by the heated plates and forced convection, due to the fan.

Figure 10 : Different positions of the fan in case of heated ceiling

Table 12 Awbi and Hatton (2000) Convective heat transfer coefficient correlations for
forced convection
Surface type Ventilation hcf hcf/hcn
regime
Walls ⎡ 0.873 ⎤
3.79(W ) ⎢2.3165(W )
1.536 1.536 U
U 0.873 ⎥
( 9 x 108 < 6 x 1010) ⎣⎢ ΔT 0.293 ⎦⎥

Floors Jet over a heated ⎡ 0.557 ⎤


4.248(W ) ⎢2.06(W )
0.575 0.595 U
U0.557 ⎥
( 9 x 108 < 7x 1010) surface ⎣⎢ ΔT 0.308 ⎦⎥
Ceilings ⎡ 0.772 ⎤
1.35(W ) ⎢3.45(W )
0.074 0.074 U
U 0.772 ⎥
( 9 x 108 < 1 x 1011) ⎣⎢ ΔT 0.133 ⎦⎥

Novoselac (2005) investigated the validity of the existing correlations from different
authors for the airflow regimes in buildings. Afterwards, he developed new convection
correlations for surface types and airflow regimes where validation of the existing
correlations failed by experimental measurements. The measurements were conducted
in an experimental chamber with typical room size (6 x 4 x 2.7 m) and typical positions
for diffusers and radiant panels (Figure 11). Adjacent to the environmental chamber is a
climate chamber to simulate external conditions. For the correlations developed with a
displacement ventilation system, air was supplied by displacement diffusers. For the
forced convection correlations with mixing ventilation systems, a high aspiration diffuser,
located at the ceiling, discharge jets along the long side of the radiant panels. The
cooling panels occupied 50% of ceiling space and they were integrated into the
suspended ceiling structure.

Figure 11 Experimental facility for the development of convection correlations


(Novoselac 2005)

2.3.4 Flow inside buildings (CFD data)

Awbi (1998) compared experimental results for natural CHTC of heated room surfaces
with CFD calculations. Two turbulence models were used: (1) a standard k- model
using wall functions and (2) a low Reynolds number k-ε model.
The logarithmic standard wall functions describe the momentum and heat transfer from
the internal surfaces of a room. But these functions are empirically derived for forced
convection in pipes and over flat plates. Awbi (1998) concluded that prediction of the
convective heat transfer coefficient using wall functions is extremely sensitive to the
distance of the point from the surface (yp) at which the wall function is applied. But CFD
analysis, which uses wall functions, proved to be useful in the investigation of the airflow
over the heated plates and the air movement within the chamber (Awbi (1998)).The
more accurate prediction of the heat transfer from room surfaces, using a low Reynolds
number turbulence model, is very time consuming.

An alternative is to use an experimental determined expression for the convective heat


transfer coefficient for room surfaces in a CFD code. This is what Beausoleil-Morrison
(2000) and Novoselac (2005) have done with their ACA and respectively MACA
algorithms.

During the Annex 41 CFD was used to evaluate the possibilities of determining heat
transfer coefficients with CFD. In [A41-T3-C-06-5 and A41-T3-C-06-6] CFD was used to
determine the heat transfer coefficient for flow between two infinite plates. A good
agreement was found in laminar flow between analytical solutions for different cases and
the CFD results if the bulk fluid temperature was used as a reference temperature (error
< 10-2 %). For turbulent forced convection a good agreement was also found and
different turbulence models give limited deviations in the developed zone. For a natural
convection case the CFD calculations did not result in velocity profiles consistent with
experimental data when using law-of-the-wall equations for forced convection. Using a
Low-Re model resulted in good agreement between experiment and simulation.

2.4 MASS TRANSFER

2.4.1 Heat and mass transfer analogy


As little direct experimental data is available on vapour transfer coefficients the heat and
mass analogy is used to calculate the vapour transfer coefficient. This is the basis for the
prEN 15061, 2004. The heat and mass transfer analogy is applicable for very specific
cases.

In case both Pr and Sc numbers are equal to one, and hence the Lewis number (Le) is
one, the relation between the mass transfer coefficient and the heat transfer coefficient
is given by :

h
= hm
ρv∞ c p

The Reynolds analogy is limited in its application because of the strict conditions under
which it is valid. (It was deduced for laminar forced flow across a solid boundary under
with no form drag.) Yet this analogy inspired researchers to seek for better analogies
which are more generally applicable. Prandtl developed an analogy for heat and
momentum transfer and for mass and momentum transfer considering the turbulent core
and the laminar sublayer in the boundary layer equations. The effect of Pr and Sc
numbers different from one is taken into account in this analogy. This led to the following
equations for the heat and mass transfer coefficients:

cf cf
h hm
St = = 2
= 2

ρv∞ c p 1 + 5 cf
S (Pr − 1) v∞ 1 + 5 cf
S (Sc − 1)
Von Karman extended Prandtl’s work and took the effect of the transition layer between
the laminar sublayer and the turbulent core into account. This led to an extra correction
term as function of respectively Pr and Sc in the two previous equations. The application
of the Prandtl and Von Karman analogies is restricted to cases with negligible form drag.
Both the Prandtl analogy as the Von Karman analogy reduce to the Reynolds analogy
for Pr and Sc number equal to one.
While Prandtl and Von Karman adapted the Reynolds analogy by considering the
transfer equations in the boundary layer, Chilton and Colburn sought modifications to the
Reynolds analogy using experimental data (Colburn 1933, Chilton & Colburn 1934).
They suggested a simple modification for situations with Pr and Sc numbers different
from unity. This was done by defining the j factor for heat transfer and the j factor for
mass transfer:

cf hm 2 / 3 c f
j h = St Pr 2 / 3 = jm = Pr =
2 v∞ 2
Colburn applied the j factor for heat transfer to a wide range of data for flow on different
geometries and found it to be quite accurate for conditions where no form drag exists
and for Pr between 0.5 and 50. The complete Chilton-Colburn analogy is found when
equations are combined:

h
Pr 2 / 3 = hm Sc 2 / 3 -
ρc p
When form drag is present neither jH or jm equals cf/2, yet it has been found that equation
remains valid. It is clear that the Chilton-Colburn analogy also reduces to the Reynolds
analogy for Pr and Sc numbers equal to unity. Unlike the Prandtl or Von Karman analogy
the relation between the heat and mass transfer coefficients is no longer function of the
skin friction coefficient.
The analogy between heat and momentum transfer and between mass and momentum
transfer is based on the assumption that respectively the dimensionless velocity and
temperature profiles and the dimensionless velocity and mass concentration profiles are
similar. This is the case for forced convection flow over a solid surface without form drag.
All the analogies mentioned in this paragraph, as well the theoretical deduced ones as
the experimental Chilton-Colburn analogy, were developed for this case.

In A41-T3-C-04-7 D. Derome presented a limited and preliminary set of experiments


through which the author suggests that the analogy between heat and mass transfer
gives an overprediction of the mass transfer. The author claims more research is needed
into the use of the analogy.

In A41-T3-B-07-4 Steeman et al. investigated the influence of conditions in which the


analogy is valid and determined accuracy. CFD simulations are performed to compare
simulated vapour transfer coefficients with coefficients predicted out of the heat and
mass analogy. It is found that dissimilarity of the boundary conditions induces the largest
differences between predicted and simulated transfer coefficients: for the average
transfer coefficient the largest difference seen in the simulations is an under prediction
with a factor 0.41 while for local transfer coefficients differences up to a factor 10 are
found. Hence the heat and mass analogy gives a reasonable estimate of the average
vapour transfer coefficient, but can lead to large errors for local coefficients.

2.4.2 Flow over and around buildings (experimental data)

Little information can be found in literature about experimentally determined mass


transfer coefficients for building applications. Swartz (1972) presented some
experimental results for mass transfer to building surfaces under different wind speeds.
Swartz reported for Vwind > 1m/s

hm = 27Vwind
0.58
windward face
hm = 20Vwind
0.43
leeward face

Worch (2004) published data for vapour transfer resistance for flows in natural en forced
convection, in and around buildings. He determined vapour transfer resistances which
are related to the mass transfer coefficient as :
′ δ
sd =
hm with δ the vapour permeability coefficient (kg/m²sPa)

Table 13 gives an overview.

Table 13 : vapour resistances according the Worch 2004.

2.4.3 Flow over building materials (numerical data)


Zhang and Niu (2003) investigated for low Reynolds numbers the validity of using CFD
for determining the mass transfer coefficients. They performed experiments on a very
small test cell (The field and laboratory emission cell (FLEC)). The authors showed that
the numerical and experimental results correlated well for different test cases.

The study of Kaya et al (2007) deals with simultaneous heat and mass transfer during
drying of cylindrical moist objects through an implicit finite-difference method.
Instantaneous temperature and moisture distributions inside the moist material as well
as all local convective heat and mass transfer coefficients are also studied via the Fluent
computational fluid dynamics (CFD) package. It is found that the convective heat transfer
coefficients vary from 4.65 to 59.33 W/m2K, while the convective mass transfer
coefficients range between 3.59 3 10-7 and 4.58 3 106 m/s, respectively. Remarkably
good agreement is obtained between the predicted results and experimental data taken
from the literature to validate the present model.
2.4.4 Flow over building materials (experimental data)

Tremblay et al (2000) measured heat and mass transfer coefficients over a flat piece of
wood inserted in a duct. They showed that the measured data correlated well with the
analytical solution for a turbulent duct flow
Sh = 0.023Sc 0.3 Re 0.8

During the Annex 41 several papers were presented in which experiments were
presented to determine the mass transfer coefficient.

Hedegaard et al (A41-T3-Dk-05-4) proposed a modified ‘cup method’ to derive a mass


transfer resistance Z (m²sPa/kg). To perform the cup method measurements specially
developed equipment has been used. The cup test facility consists of a closed
ventilation system where both temperature and RH can be controlled. A diagram of the
equipment can be seen in Figure 12 (1).

Figure 12: Test setup of Hedegaard et al (A41-T3-Dk-05-4)

During the test the temperature, RH, airflow velocity and pressure is recorded
automatically and the weight of the cups is entered at each weighing. It is possible to
test 12 ordinary cups and 12 inverted cups at the same time. In Figure 12 (2) a picture of
the test chamber is shown. In the picture two holes can be seen in the front plate and by
use of these the samples are weighed on the balance seen in the bottom. A more
detailed description of the used equipment is given by Hansen (1989). The air is
circulated by a fan in a squared duct, which is stretched out to a flat 60 cm wide duct of 5
cm height in the lower part of the system. In this flat part of the duct the cup samples are
in contact with the chamber air. The circulation ensures that the airflow velocity on the
exterior side of the cups can be controlled. Different airflow rates can be set.
Different materials were tested under different conditions as shown in table 2
Table 14 : Samples tested by Hedegaard et al (A41-T3-Dk-05-4)

The results of the tests A1-A5 do not clearly indicate how the air flow influences the total
resistance of the material sample. It was impossible to conclude anything about the
influence of the airflow velocities influence on the surface resistances.
The results of the tests B1-B4 with the glass fibre membranes were somewhat ruined
because the salt solution in the wet cup crept up at the sides of the cups and onto the
material samples. In the cases with most salt on the samples the material resistance was
highly reduced. Therefore, the tests of the surface resistances with wet cups were
abandoned.
The results shown for the tests C1-C4 consists of a number of measurements with
different airflow velocities based on the measuring results from each cup. An example of
the weight uptake results for the tests with paper are given in Figure 13 for the dry cups
in test C1. In the figure the numbers in the legend refer to the airflow velocity above the
boundary layer of the material surface of the given cup. The slopes of the lines in
combination with the exposed surface area and the vapor water pressure difference over
the samples are used to calculate the total resistance of the samples. The slopes of the
lines in Figure 13 are for 2-layers of paper. The calculated total resistances in this case
are between 8.96·107 and 1.23·108 Pa m2 s/ kg. The lowest surface resistance is for the
case with a velocity of 0.34 m/s and the highest value corresponds to an airflow velocity
of 0.06 m/s. These airflow velocities also have the steepest and the flattest slopes
respectively in Figure 13. The measured weight uptake rates have been post processed
and the corresponding surface coefficients have been found. The corresponding surface
resistances as a function of the airflow velocity above the boundary layer of the four tests
C1-C4 are shown in Figure 14. The results shown in the figure are based on at least 5
weightings where the weight change rate is constant within ± 5% of the mean value,
which is required by the EN ISO 12572:2001 standard. However, in most cases the
weight change rate was constant within ± 2% of the mean value. In the figure a trendline
calculated by the least squares fit for all measured test results by use of a power function
are added.
The measured results in Figure 14 show that there is a tendency of higher surface
resistances for lower airflow velocities. This was expected. However, if the surface
resistances from the measurements are compared with the estimated surface
resistances, it is found that the measured values are higher than predicted. This could
be a sign of that the equations slightly underestimate the surface resistances. For
comparison the results of Bednar & Dreyer (2003) showed that the moisture transfer
coefficient for drying is around 18·10-05 kg/(h m2 Pa) for a room with .still. air. This
number can be converted to a surface resistance value of 2.0·107 Pa m2 s/ kg . This
number seems quite small compared to measurement results where both the estimated
and the measured surface resistances for velocities less than 0.2 m/s are higher.
However, it was a drying experiment where the sample was wet and since the liquid
mass transfer within the sample is faster than evaporation this can explain the lower
value. In the present study the difference between the estimated values by use of Lewis
relation and the measured values decreases as the airflow velocity is increased.
However, the normal airflow velocities in dwellings near construction surfaces are often
quite small so there the underestimated values could be a problem.

Figure 13 : Test results of Hedegaard et al (A41-T3-Dk-05-4)


Fi

Figure 14 : Test results of Hedegaard et al (A41-T3-Dk-05-4)

Talev et al (A41-T3-N-06-2) explore the role of transport properties of moist air as well as
the air velocity on the convective surface mass transfer coefficients at different axial
positions in a rectangular cross-section wind tunnel. Experimental work has been
performed to determine the local mass transfer coefficients using three equal, horizontal
water cups, placed inline after one another in the tunnel. Each of the three samples
holders had a square shape with length and width equal to 60 mm and was mounted in
line with the bottom surface of the wind tunnel, so that the air stream passed over the
water surface (Figure 15).
Figure 15 : Test setup of Talev et al (A41-T3-N-06-2)

A series of experiments was performed to determine the resistance of the moisture


transport between the free water surface and moist air as a function of air velocity,
distance from the tunnel opening and the relative humidity (RH). All the experiments
were carried out for a moist air temperature of 20 °C. Some results are shown in Figure
16. Each figure shows the surface mass transfer coefficient on the vertical axis in units
kg/ (Pa·s ·m2). The horizontal axis shows the airflow velocity at the tunnel entrance.
Further, each diagram includes results from 5 measurements, noted Measurement 1 to
5, and the average of these measurements (noted “Average Measurement”). In addition
the figures include results from correlations found in literature which are noted “Theory 1
Laminar”, “Theory 1 Turbulent”, and “Theory 2”, respectively. Notice that “Theory 2” is
intended for airflow velocity lower than 5 m/s.
First the results in the figures show that there is some spread in the experimental data.
This is mainly attributed by the fact that there was very difficult to maintain a perfectly flat
water surface. The measurements showed that a convex surface had a higher mass
transfer coefficient than a flat surface. A concave water surface has a lower mass
transfer coefficient than a flat water surface (this can not be expressed by the
correlations presented in this paper). The water supply system was also quite difficult to
control. Still, the measured results show the same trend. Questions could also be raised
about whether the measured results represent an ideal external flow or not. For a
position of 37 cm (for cup 1) and 61 cm (for cup3) from the tunnel entrance the velocity
boundary layer thickness will be about 4 cm and 5 cm (calculated with an equation from
White, 1999), respectively, at a velocity of 0,1 m/s. At a free stream velocity of 1 m/s the
boundary layer thickness will be about 1, 2 cm for cup 1 and 1,5 cm for cup 3. Generally,
a large Reynolds number represents a thinner boundary layer (at a fixed position, x),
while a small Reynolds number represents in a thicker boundary layer. Thus, data for
low velocities may therefore not be representative for a real external flow. (Later,
experiments will be carried out in a modified wind tunnel to ensure the results are for
external flow, for all velocities.
The figures show that the surface mass transfer coefficient was a function of the airflow
velocity ,local position and the relative humidity All figures show that the surface
moisture transfer coefficient increases with the air velocity. This is because an increased
velocity reduced the boundary layer thickness. The measurements show that an
increase in the RH resulted in a decrease in the surface mass transfer coefficients. An
increased distance from the tunnel entrance resulted in a reduction of the surface
moisture transfer coefficient.
Figure 16 : Test results of Talev et al (A41-T3-N-06-2)

In general, there is quite poor agreement between the measured and the theoretical
results. Lack of resemblance between the measured data and theoretical equations may
be because the theoretical equations do not take into account all the processes taking
place (e.g. thermal radiation, ‘blowing’ effects (non-zero transverse velocity at interface),
etc.). Further analysis is required in order to explain the observed discrepancies.

C Iskra and CJ. Simonson (A41-T3-C-06-3) (See also Iskra and Simonson CJ (2007)
and ANNEX 41 Subtask 2; 3.3) measured the convective mass transfer coefficient
between a forced convection airflow and a free water surface using the transient
moisture transfer (TMT) facility at the University of Saskatchewan. A pan of water is
situated in the test section and forms the lower panel of the rectangular duct, where a
hydrodynamically fully developed laminar or turbulent airflow is passed over the surface
of the water. As the air passes through the test section, it loses heat and gains moisture
to/from the water. As a result, the thermal and concentrations boundary layers are
developing through the short test section. The experimental data shows that the
convective mass transfer coefficient is a function of the Reynolds number (570 < Re <
8,100 investigated) and the relative humidity of the air stream (15% to 80% RH
investigated). The air humidity can change the mass transfer coefficient by as much as
35%. For example with Re = 1500, the measured Sherwood number is 5.1 and 6.9 (hm =
0.0031 and 0.0042 m/s) when the air humidity is 80% RH and 18% RH respectively.
The transient moisture transport (TMT) facility is an experimental apparatus that
determines the transient heat and moisture transport properties of porous materials. The
test section within the TMT is a horizontal rectangular duct, where heat and mass
transfer occurs at the bottom wall surface. The facility passes air at varying velocities,
temperatures and relative humidity’s above the surface of materials and measures the
change in mass, relative humidity, and temperature in the material as a function of time.
To fully document experimental results for the TMT, the convective mass transfer
coefficient of this facility is required Mass transfer coefficients are usually determined
from experiments based on the adiabatic evaporation of a liquid and this method is
applied in this paper. Since water vapor transfer is the only form of mass transfer in the
TMT facility when experimenting on porous materials, distilled water is used as the
evaporating liquid to ensure that the same Schmidt number (Sc) is present in both
experiments. The convective mass transfer coefficient is determined for the horizontal
rectangular duct by measuring the evaporation rate from a rectangular tray of water that
is located in the lower panel of the duct. The vapor density difference between the bulk
air stream and the surface of the water is also measured in order to determine the
concentration difference at the surface of the duct.
A side-view and an expanded top-view schematic of the ducting upstream and
downstream of the rectangular test section are shown in Fig. 17(a and b). All of the
rectangular ducting shown have a width (W) of 298 mm. A variable speed vacuum pump
supplies a hydrodynamically developed airflow at the entrance of the test section by
means of a developing section upstream of the test section (Fig. 17(a)). First, the air
flows through a 1100 mm long duct that has a constant cross sectional area, which has
several screens installed inside of it to aid in the straightening of the airflow. The air then
passes through a 995 mm long converging section that has a convergence angle of 5° to
minimize the dynamic losses in the duct and aid in the development of the flow.
Following the converging section, the air enters a 500 mm long straight duct that delivers
the air to the test section. The air then passes through the 765 mm test section (close-up
view shown in Fig. 17(c)) and then through a 890 mm downstream section. The ducts
immediately upstream and downstream of the test section have the same hydraulic
diameter as the test section when the evaporation pan is full of water (i.e., h = 0 in Fig.
17(d)).
The side and top views of the test section within the TMT facility are also shown in Fig.
1. A tray with a water surface width (w) of 280 mm and a length (L) of 600 mm forms the
lower panel of the duct (height (H) = 20.5 mm) in the test section. The air is delivered to
the test section from an environmental chamber that controls the temperature and
relative humidity of the air upstream of the test section within ± 0.1°C and ± 2% RH
respectively. The temperature and relative humidity of the air is measured upstream and
downstream of the test section with humidity/temperature transmitters.
Figure 17 : TMT test setup of Iskra and Simonson (A41-T3-C-06-3)

Laminar flow

Figure 18 presents the convective mass transfer coefficients, hm (m/s) for a range of Re
and air relative humidity. The convective mass transfer coefficients decrease when the
relative humidity of the air entering the test section is increased. The convective mass
transfer coefficients can change as much as 35% when the air relative humidity varies
from 18 to 80% RH. To fully capture the effects of the relative humidity on the convective
mass transfer coefficient, the Rayleigh number (Ra=Re Pr) is used, which takes into
account the temperature and relative humidity of the air at the surface of the water and
in the bulk airflow. The convective mass transfer coefficient (hm) is non-dimensionalized
with the use of Sh and this data is presented in Fig. 19.
The Sh is determined for laminar flow between a Re of 570 and 2,100, which
corresponds to a X* between 0.011 and 0.037, and a Ra of 6,300 and 83,000. A general
trend of increasing Sh as Re increases (X* decreases),is found which is expected since
the length of the test section is not long enough for the temperature and concentration
boundary layers to become fully developed. As Re increases, the thermal and
concentration boundary layers become thinner and less developed over the surface of
the water. A thinner boundary layer results in a larger concentration gradient at the
surface of the water, which contributes to an increase in forced convection mass transfer
at the surface of the water.

with

Figure 18 : Mass transfer coefficient as function of Re en relative humidity Iskra and


Simonson (A41-T3-C-06-3)
(
Figure 19 : Sh as function of Re en relative humidity Iskra and Simonson (A41-T3-C-06-
3)

Turbulent flow

Turbulent flow experiments are performed in order to develop a relationship for Sh that
includes both developing flow (X*) and buoyancy forces (Ra), and also to further verify
the experiments by comparing to an experiment in the literature which is for turbulent
flow. The turbulent flow data of the present work covers a range of Re between 3,100
and 8,100, and Ra between 20,900 and 46,000.
The various Ra are created by the air relative humidity between 15% RH and 60% RH at
a constant air temperature of 23°C. The results that as Ra increases, Sh increases. The
contribution of natural convection compared to that of forced convection is measured by
Ri=Gr/Re². It is found that this ratio is approximately 20 times less than that for the
laminar flow experiments, which suggests that the contribution of natural convection
evaporation is smaller in the presence of forced convection turbulent flow than in laminar
flow. Nevertheless, the effects of Ra exists and the experimental data in the turbulent
region are correlated with X* and Ra, which results in

2.5 CONCLUSION

Transfer coefficients for heat and mass transfer between air flows and building surfaces
are the result of modelling assumptions under the boundary layer theory.
For heat transfer coefficients a lot of experimental data is available to determine the
convection transfer coefficient, both for natural and forced convection. Care must be
taken when selecting the appropriate geometry and flow regime. Recently CFD
simulations are also used for determining the heat transfer coefficient. The first results
show that the accuracy is good. Further validation is needed.
Mass transfer coefficients are not readily determined experimentally. Little information
can be found in literature. Mostly the heat transfer coefficient is determined and mass
transfer is calculated with the heat and mass transfer analogy. Again, and even more,
care must be taken when selecting the appropriate geometry and flow regime.
New experimental methods were proposed during the annex but still a lot of
development has to be done.
REFERENCES

ASHRAE (1975) Task Group, Procedure for determining heating and cooling loads for
computerising energy calculations, in: Algorithms for Building Heat Transfer Subroutines,
ASHRAE Publications, New York.

Alamdari F. and Hammond G.P. (1983). Improved data correlations for buoyancy-driven
convection in rooms. Building Services Engineering Research and Technology 4 (3):
106-112.

Awbi H.B. and Hatton A. (1999). Natural convection from heated room surfaces. Energy
and Buildings 30(3): 233-244.

Awbi H.B. and Hatton A. (2000). Mixed convection from heated room surfaces. Energy
and Buildings 32(2): 153-166.

Beausoleil-Morrison I. (1999). Modelling mixed convection heat transfer at internal


building surfaces. Proc. International Building Performance Simulation Association.
Kyoto, Japan.

Beausoleil-Morrison I. (2000). The adaptive coupling of heat and air flow modelling within
dynamic whole-building simulation. Ph.D. thesis, University of Strathclyde, Glasgow, UK.

Bednar, T. & Dreyer, J. (2003). Determination of moisture surface transfer coefficients


under transient conditions. Proceedings of the 2nd International Conference on Building
Physics, Leuven, Belgium, p.233-236. ISBN 90 5809 565 7.

Chilton, T.H. and Colburn. A.P., (1934), Mass Transfer (Absorption) Coefficients
Prediction from Data on Heat Transfer and Fluid Friction. Industrial and Engineering
Chemistry,. 26: 1183-1187.

CIBSE (1986). CIBSE Guide, Vols. A, B and C, CIBSE, London.

Clarke J.A. (2001) Energy simulation in building design. 2nd Edition. Oxford,
Butterworth-Heinemann.

Colburn, A.P., (1933), A method of correlating forced convection heat transfer data and
a comparison with fluid friction. Transactions of the American Institute of Chemical
Engineers,. 29: 174-210.

Fisher D.E. (1995). An experimental investigation of mixed convection heat transfer in a


rectangular enclosure. Ph.D. thesis, University of Illinois, Urbana, USA.

Fisher D.E. and Pedersen C.O. (1997). Convective heat transfer in building energy and
thermal load calculations. ASHRAE Transactions 103 (2): 137-148.

Hansen, K.(1989). Equipment for and results of water vapour transmission tests using
cup methods. Proceedings from the ICHMT Symposium .Heat and Mass Transfer in
Building Materials and Structures.. Dubrovnik, Yugoslavia.
Iskra CR, Simonson CJ (2007) Convective mass transfer coefficient for a
hydrodynamically developed airflow in a short rectangular duct. International Journal of
Heat and Mass Transfer 50: 2376-2393

Judkoff R.D., Neymark J.S. (1995), Building Energy Simulation Test (BESTEST) and
Diagnostic Method, NREL/TP-472-6231, Golden, Colorado National Renewable Energy
Laboratory.

Kaya A, Aydin O, Dincer I (2007) Numerical modeling of heat and mass transfer during
forced convection drying of rectangular moist objects. International Journal of Heat and
Mass Transfer 49: 3094-3103

Khalifa A.J.N. (1989). Heat transfer processes in buildings. Ph.D. Thesis, University of
Wales College of Cardiff, Cardiff, UK.

Khalifa A.J.N. (2001). Natural convective heat transfer coefficient – a review: I. Isolated
vertical and horizontal surfaces. Energy Conversion and Management 42(4): 491-504.

Khalifa A.J.N. (2001). Natural convective heat transfer coefficient – a review: II. Surfaces
in two- and three-dimensional enclosures. Energy Conversion and Management 42(4):
505-517.

Khalifa A.J.N. and M. R.H. (1990). Validation of heat transfer coefficients on interior
building surfaces using a real-sized indoor test cell. International Journal of Heat and
Mass Transfer 33(10): 2219-2236.

Kays WM, Crawford ME (1993), Convective Heat and Mass Transfer, McGraw-Hill.

Emmel M, Abadie M, Mendes N (2007) New external convective heat transfer coefficient
correlations for isolated low-rise buildings Energy and Buildings 39 335–342

Min T., Schutrum L., et al. (1956). Natural convection and radiation in a panel heated
room. ASHRAE Transactions 62: 337-358.

Novoselac A. (2005). Combined airflow and energy simulation program for building
mechanical system design. Ph.D. thesis, Pennsylvania State University, Pennsylvania,
USA.

prEN 15026, 2004, Hygrothermal performance of building components and building


elements – Assessment of moisture transfer by numerical simulation.
DRAFT August 2007
IEA ANNEX 41 MOISTURE ENGINEERING SUBTASK 3
BOUNDARY CONDITIONS AND WHOLE BUILDING HAM ANALYSIS
Final Report M De Paepe

Spitler J.D., Pedersen C.O., et al. (1991a). An experimental facility for investigation of
interior convective heat transfer. ASHRAE Transactions 97(1): 497-504.

Spitler J.D., Pedersen C.O., et al. (1991b). Interior convective heat transfer in buildings
with large ventilative flow rates. ASHRAE Transactions 97(1): 505-515.

Swartz B. (1972) Die Wärme- und Stoffübertragung an Ausswandoberflächen


Dissertation des Verfassers an der Universität Stuttgart, Institut für Bauphysic
Fraunhofer-Geselschaft Aussenstelle Holzkirchen.

Tremblay C, Cloutier A, Fortin Y (2000) Experimental determination of the convective


heat and mass transfer coefficients for wood drying. Wood Science and Technology 34:
253-276

Wadso L (1993) Surface Mass-Transfer Coefficients for Wood. Drying Technology 11:
1227-1249

Welty JR, Wicks CE, Wilson RE, Rorrer G (2001), Fundamentals of Momentum, Heat
and Mass Transfer, John Wiley & Sons.

Worch A (2004) The Behaviour of Vapour Transfer on Building Material Surfaces: The
Vapour Transfer Resistance. Journal of Thermal Envelope and Building Science 28:
187-200 DOI 10.1177/1097196304044398

Zhang LZ, Niu JL (2003) Laminar fluid flow and mass transfer in a standard field and
laboratory emission cell. International Journal of Heat and Mass Transfer 46: 91-100

Papers presented during ANNEX 41

A41-T3-C-04-7 D. Derome, Experimental determination of the convective mass transfer


coefficient

A41-T3-Dk-05-4 L Hedegaard Mortensen, C. Rode, R Peuhkuri Effect of airflow velocity


on moisture exchange at surfaces

A41-T3-N-06-2 G Talev, A Gustavsen and E Næss The influence of air velocity and
transport properties on the surface mass transfer coefficient in a rectangular tunnel –
theory and experiments

A41-T3-C-06-3 C Iskra, CJ. Simonson Effect of air humidity on the convective mass
transfer coefficient in a rectangular duct

A41-T3-Br-07-2 Marcelo G. Emmel, Marc O. Abadie, Nathan Mendes New external


convective heat transfer coefficient correlations for isolated low-rise buildings

A41-T3-B-07-4 H.-J. Steeman, A. Janssens, M. De Paepe About the use of the heat and
mass analogy in building simulation

37

Das könnte Ihnen auch gefallen