You are on page 1of 15

Applied Thermal Engineering 87 (2015) 258e272

Contents lists available at ScienceDirect

Applied Thermal Engineering


journal homepage: www.elsevier.com/locate/apthermeng

Research paper

Variability and impact of internal surfaces convective heat transfer


coefficients in the thermal evaluation of office buildings
Sophie Obyn*, Geoffrey van Moeseke 1
Architecture et Climat, Facult
e d'architecture, d'ing
enierie architecturale, d'urbanisme (LOCI), Universit
e Catholique de Louvain, 1 place du Levant,
1348 Louvain-la-Neuve, Belgium

h i g h l i g h t s

 Despite significant deviations of CHTC values, those of energetic demands are limited.
 CHTC expression slightly impacts HVAC installation design for light structures.
 For heavy structures, CHTC expression has a major impact on HVAC installation design.
 Constant value of CHTC may be considered to evaluate energetic loads and demands.
 Expressions of category 4 do not seem to be appropriate in the construction sector.

a r t i c l e i n f o a b s t r a c t

Article history: The choice of surface convective heat transfer coefficient (CHTC) affects the evaluation of buildings'
Received 5 January 2015 thermal comfort and energy consumption. There are many established analytical and experimental
Accepted 13 May 2015 expressions of these coefficients for different flow types and ventilation modes. The aim of the present
Available online 22 May 2015
work is to investigate the impact of CHTC expressions used for the calculation of heating and cooling
loads of a building. Therefore, a compilation and a classification of CHTC expressions into four categories
Keywords:
are proposed. Then, these correlations are applied on the case of a parameterized office room, consid-
Heat transfer
ering one type of surface at a time (wall, ceiling, floor).
Convection
Convection coefficient
As conclusions, despite significant deviations of CHTC values calculated, the deviations of energetic
Convection correlation demands are limited at the room scale. Also, the impact of CHTC expression on HVAC installation design
Energy consumption is quite limited for light structures but is major for heavy structures. Otherwise, the use of constant CHTC
Buildings value appears to be acceptable in many cases, despite the fact that deviations in heating demand of up to
Sensitivity analysis 9.1% occur in some cases, compared to temperature-dependent expression. Finally, the use of more
Simulation complex CHTC expressions (referred to as “mixed expressions” in this article) does not seem to be
Building valuable in the construction sector.
Modeling
© 2015 Elsevier Ltd. All rights reserved.
Critical review

1. Introduction other the CHTC. Also, previous studies provide different values and
expressions to evaluation this CHTC.
This research is based on two main observations. Firstly, build- The aim of this study is to evaluate the impact of various ex-
ing energy simulation programs are important tools in building pressions of CHTC evaluating heating and cooling demands and
design and operation [1e4] and they are increasingly used, among maximal loads for a standard office building. There are different
other by consulting and engineering offices. On the other hand, methods to obtain values for CHTC which can be categorized in
these programs consider a large amount of assumptions, among analytical, numerical and experimental methods, as developed by
Mirsadeghi et al. in 2013 for external CHTC [5]. The expressions of
CHTC considered in this study (listed in Appendix A) are described
* Corresponding author. Tel.: þ32 10 47 21 42; fax: þ32 10 47 21 50. and a classification is proposed in the first subchapter of this
E-mail addresses: sophie.obyn@uclouvain.be (S. Obyn), geoffrey.vanmoeseke@ introduction. The second subchapter of the introduction is a review
uclouvain.be (G. van Moeseke).
1 of the origins and usages of these correlations.
Tel.: þ32 10 47 21 42; fax: þ32 10 47 21 50.

http://dx.doi.org/10.1016/j.applthermaleng.2015.05.030
1359-4311/© 2015 Elsevier Ltd. All rights reserved.
S. Obyn, G. van Moeseke / Applied Thermal Engineering 87 (2015) 258e272 259

Nomenclature RaH* ¼ g b Q H4/(n k a) modified Rayleigh number based on H [/]


Ree ¼ r$V_
1=3
=m$V
diffuser enclosure Reynolds number [/]
room
A area of the considered surface [m2] Tair average temperature of the ambient air [K]
ACH ventilation rate measured in room air changes per hour Tref reference (air) temperature [K]
[h1] Tsurf Average temperature of the surface [K]
CHTC convective heat transfer coefficient [W/(m2 K)] Tdiffuser supply air temperature at the diffuser [K]
Cp specific heat capacity [kJ/(kg K)]
V_ diffuser volumetric flow rate of air entering the room at the
D characteristic dimension [m]
Dh ¼ 4 A/Perimeter hydraulic diameter of considered surface [m] diffuser [m3/s]
Fo Fourier number [/] Vroom room volume [m3]
g gravitational acceleration [m/s2] WWR window to wall ratio
Gr ¼ g b DT D3h/n2 average Grashof number [/]
hc average convective heat transfer coefficient [W/(m2 K)] Greek letters
H height of vertical surface [m] a thermal diffusivity [m2/s]
k thermal conductivity of air [W/(m K)] b volume coefficient of expansion [K1]
L length of the room [m] DT ¼ jTsurface  Tairj average temperature difference [K]
length length of external wall with glazing in the zone [m] DTvent ¼ jTair_supply  Tair_extractj absolute temperature difference
Nu ¼ hc L/k Nusselt number [/] [K]
Pr ¼ ʋ/a Prandtl number [/] m molecular dynamic viscosity of air [Pa s]
Q heat flux density [W/m2] r air density [kg/m3]
qc convective heat flux density [W/m2] n kinematic viscosity [m2/s]
Ra ¼ Cp r g b DT D3/(n k) average Rayleigh number [/]

1.1. Expressions of CHTC Figs. 1 and 2 illustrate the ranges of values provided by the
different correlation categories. It immediately appears that cor-
The CHTC, expressed as hc, is defined by Eq. (1), where qc is the relations of category 4 are functions of (1/DT), which is a priori not
convective heat flux density, Tsurf is the surface temperature, and suitable for building simulations which deal with small tempera-
Tref is the reference (air) temperature. Obviously CHTC varies with ture differences. Fig. 1 highlights large differences in terms of CHTC
the temperature difference between surface and ambient air. value for the same category of expressions. Indeed, for a tempera-
However, since the value of qc and the temperature difference are ture difference of 2  C, the CHTC value may vary from 0.5 to 4.7 W/
the desired outputs in building thermal models, hc is estimated as a (m2 K) for vertical surface correlations and may rise up to 8e9 W/
function of the Nusselt number using Eq. (2), where Nu is a function (m2 K) with specific correlations for windows. Most expressions of
of Ra and Pr for natural convection and a function of Re, Pr, and L/D category 3 (hc ¼ f(airflow rate)) provide values well below the other
for forced convection. Therefore, CHTC is a function of surface ge- categories for the range of airflow typical of an office space (Fig. 2):
ometry, air properties, and temperature difference between surface 0.2e1 W/(m2 K) for the majority and 1.1e8.1 W/(m2 K) for Fisher
and ambient air. and Pedersen correlations [27].
The convection mode (natural, mixed or forced) corresponding
qc h . i to each CHTC expression is not always specified. Nevertheless, the
hc ¼  
 W m2 K (1)
ranges of CHTC values obtained for each mode are very larges and
Tsurf  Tref 
overlap. Moreover, each convection mode can be met in a building
in function of heating and ventilation modes considered, as speci-
Nu*k h . i fied by Awbi and Hatton in 200 [31]. Consequently, expressions
hc ¼ W m2 K (2) corresponding of all convection modes are considered in this study.
L
Based on a bibliographic survey and a compilation of CHTC ex- 1.2. Review of the literature
pressions available on the literature, a classification of CHTC ex-
pressions into four categories is proposed: (1) hc as a constant Papers on which this study is relied are presented in this sub-
value; (2) hc as a function of the temperature difference between chapter, in chronological order of publication.
the surface and inside temperatures of the room; (3) hc as a func- In 1983 Alamdari and Hammond [16] proposed correlations of
tion of the air exchange rate in the building or type of airflow in the CHTC for internal surfaces in naturally ventilated buildings based
room (laminar or turbulent); and (4) mixed expressions. All on the mathematical model developed by Churchill and Usagi [32]
recorded equations are shown in Appendix A. Table 1 summarizes in 1972. These expressions correspond to categories 2.1 and 2.3
the proposed categories of CHTC expressions. (hc ¼ f(DT)) and cover the full range of laminar, transitional, and
There also exist expressions related to the width of the nozzle turbulent airflows. These expressions compare favorably with
and the velocity of air at the nozzle in cases of forced convection, as available experimental data for isolated surfaces and are presented
proposed by Awbi and Hatton in 2000 [31]. Such expressions are as an improvement over the “standard” correlations recommended
considered inapplicable to the general study presented here in the CIBSE Guide [19]. Churchill and Usagi's formula [32] is in
because they focus on detailed HVAC technical development, while category 2.3 and presents the formulation of Eq. (3).
the present study addresses the accuracy of thermal models at the
("   #6 )
building scale. Furthermore, CHTC expressions with a range of
DT 0:25 h i6 1=6
validity that does not match the field of building physics are hcy ¼ c þ d ðDTÞ0:33 (3)
y
excluded from the review of this paper.
260 S. Obyn, G. van Moeseke / Applied Thermal Engineering 87 (2015) 258e272

Table 1
Proposed categories of CHTC expressions, defined by generic expressions.

Category of Formulation Remark Authors


expression

Category 1: Constant value


1.1. Constant value hc ¼ Constant / [6e8]

1.2. Constant value C1 if …  Condition in function of type of airflow (reduce or enhance convection) or DT > or < 0 K [6,8]
hc ¼
with condition C2 if …
Category 2: Function of the temperature difference
2.1. Temperature hc ¼ a ðDTÞb  a is sometimes a function of the geometry of the office [9e21]
difference (  b is a constant
b1
a1 ðDTÞ if …
2.2.Temperature hc ¼ b2
 type of air flow (laminar or turbulent); [9,11e14,16e19,22e24]
difference with a2 ðDTÞ if …  DT > or < 0 K;
condition  a is sometimes a function of the geometry of the office
 b is a constant
b1 c1
2.3. Complex hc ¼ ½fa1 ðDTÞ g þ fa2 ðDTÞb2 gc2 d [16]
function of
temperature
difference 
2.4. Grashof number hc ¼ a kL ðGrÞb  always function of type of air flow (laminar or turbulent) [19]
 a is a function of the geometry of the office
  b is a constant
2.5. Rayleigh number hc ¼ a k
L ðRaÞb  always function of type of air flow (laminar or turbulent) [25]
 a is always function of geometry
 b is a constant
Category 3: Function of air flow
3.1. Air change per hc ¼ a ðACHÞb þ c  a, b and c are constants [26,27]
hour
3.2. Air flow rate hc ¼ a ðV_ Þb  a is a function of the geometry of the office [28]
 b is a constant
b
3.3. Reynolds hc ¼ a ðReÞ þ c  a and c are always function of the geometry of the office [26]
number - type of  b is a constant
flow
Category 4: Mixed function
 
4. Mixed function 1 ; ACH; geometry of the office [29,30]
hc ¼ f DT; DT

Fig. 1. Range of CHTC values according to the various expressions of the proposed CHTC expressions categories, expressed as a function of temperature difference between the
considered surface and the ambient air (DT): Left  Vertical surfaces, Center  Ceiling, Right  Floor.

In 1986 Akbari et al. [33] studied the effect of variations in CHTC In 1999, Awbi and Hatton [9] proposed CHTC correlations
on thermal energy storage in structural materials in buildings. They for natural convection from heated room surfaces, based on
performed a theoretical study and detailed analytical and numer- experimental results in an environmental chamber. These expres-
ical analyses for simple geometries. In 1990 Khalifa and Marshall sions rely on the temperature difference and the geometry of the
[11] conducted experiments to determine CHTC correlations on heating surfaces. In 2000 [31], they published an article on mixed
interior building surfaces using a full-sized indoor test cell. They convection from heated room surfaces. After experiments con-
proposed 10 new expressions to be used by engineering offices. ducted in a small-office-sized, well-insulated environmental
These correspond to category 2.1 (hc ¼ f(DT)) and differ by a factor chamber, they showed that when mechanical ventilation was used,
of 2 from those commonly used in building thermal models (i.e., their former expressions of CHTC (determined in natural ventila-
Alamdari and Hammond's correlations [16]). tion conditions) would underestimate the convective heat transfer
S. Obyn, G. van Moeseke / Applied Thermal Engineering 87 (2015) 258e272 261

Watson [39]. They observed a good agreement between the


experimental correlations of Alamdari and Hammond and the
theoretical expressions they proposed. However, their theoretical
correlation led to slightly higher values of CHTC (the difference is
less than 5%), which were closer to experimental data.
In 2008, Gao et al. [40] proposed a turbulence model with a
differential viscosity and dynamic turbulent Prandtl model to pre-
dict the convective heat transfer for the case of thermally stratified
flows in buildings. They numerically derived local and bulk con-
vection coefficients based the law-of-the-wall method. Causone
et al. [41] in 2009 published a paper on experimental evaluation of
heat transfer coefficients between radiant ceiling and room. They
made experimental measurements in a test chamber to determine
values of CHTC in typical conditions of occupancy of an office or
residential building and obtained results similar to the correlations
of Khalifa and Marshall [11] (experiments on interior building
surfaces using a full-sized indoor test cell) and Khalifa [36] (review
article on natural CHTC correlations for isolated surfaces) for cooled
radiant ceilings and of Awbi and Hatton [9] (CHTC correlations for
natural convection from heated room surfaces, based on experi-
mental results in an environmental chamber) for heated ones.
In 2011, Peeters et al. published a critical review and discussion
of experimentally derived correlations of internal CHTC [42].
Leenknegt et al. [43] published a paper in 2012 on the numerical
sensitivity study of transient surface convection during night
Fig. 2. Range of CHTC values provided by correlation functions of the airflow (category cooling. They conducted a sensitivity of the influence of model
3) and the constant value provided in the TRNSYS manual [7], for an office of 48.6 m3.
simplifications on the value of CHTC and observed large variations.
However, the results were not validated by measurement and
depended on the selected turbulence model.
from the room surfaces. This underestimation varies with the sur- In 2012 Al-Sanea et al. [44] studied the effect of Reynolds
face considered (wall, floor, or ceiling). Awbi and Hatton presented number and room aspect ratio on flow and ceiling heat transfer
equations for predicting mixed convection heat transfer as a com- coefficients for mixed ventilation. They conducted a CFD study to
bination of natural and forced convection for walls, floor, and simulate turbulent flow and heat transfer inside mechanically
ceiling. These correlations are functions of the velocity of air at the ventilated rooms, using a mixing air-distribution system under
nozzle opening and of the width of the nozzle opening and provide forced convection conditions. Their numerical model was validated
higher values of CHTC than correlations for natural convection. by comparing results with available experimental data. They
Beausoleil-Morrison theorized an adaptive coupling of heat and concluded in part that ceiling-averaged convection coefficient in-
airflow modeling within dynamic whole-building simulation [29]. creases with Re and decreases with L/H.
His work was also described in a conference paper in 2001 [34] and In 2013 Baïri [45] proposed NueRaeFo correlations that can be
in two articles in 2001 [30] and 2002 [35]. He proposed a flow- used in thermal engineering applications such a building inter alea.
responsive algorithm which was implemented within the ESP-r These correlations used Fo which is difficult to define in global
simulation program to advance the modeling of internal surface annual evaluation of building energy performances. Moreover, the
convection. Empirical methods were extracted from the literature, correlations proposed look like those presented by Yousef et al. [25].
and a new method for characterizing mixed flow was created to Finally, in 2013 Ruivo et al. [46] extend the Cooling Load Tem-
provide the algorithm based on 28 CHTC correlations. All of these perature Difference (CLTD) to cooling and heating loads estimation
are studied in the present paper. Beausoleil-Morrison personally of rooms with daily and weekend setback and setup thermostats,
developed five of these 28 correlations, which correspond to introducing the term Thermal Load Temperature Difference (TLTD).
category 4 (mixed function). Therefore, they create a transient heat transfer model and the
In 2001, Khalifa [36,37] published two review articles (part I and corresponding numerical tool. Among their key assumptions, there
part II of the same review) on natural CHTC correlations for isolated is the consideration of constants combined surface heat transfer
surfaces and for two- and three-dimensional enclosures. His coefficients that account for both convection with the ambient air
conclusion was that large differences can appear between compiled and radiation with the surrounding surfaces, in both inner and
correlations. outer sides of the envelope. In 2015 Yan et al. [47] proposed a
Fohanno and Polidori [23] in 2006 proposed expressions for simplified model to evaluate the impact of radiant heat on building
both local and average CHTC in the cases of laminar and turbulent cooling load. Therefore, they developed a simplified analytical
flows on uniformly heated internal building surfaces (walls, win- model that they validated in EnergyPlus program, considering a
dows, and so forth). These expressions are derived from theoretical constant value for internal CHTC of 3.1 W/(m2 K) which is a similar
analysis based on the integral formalism. One of these is presented value than those proposed in the TRNSYS manual [7] and by ASH-
in Appendix A and corresponds to category 2.2 (function of DT). RAE in 1985 [6].
They compared their expressions with the one provided by Alam- The present work complements all these earlier works in which
dari and Hammond [16] (for internal surfaces in naturally venti- the CHTC values are considered by evaluating the impact of these
lated buildings) and experimental data on full-sized enclosures values on heating and cooling demands and maximal loads. A wider
published by Allard [38] (dwelling cells in artificial climatic con- purpose of the research is to highlight the relevance of a wise and
ditions), Awbi and Hatton [9] (for natural convection from heated conscious choice of the convection heat transfer coefficient values
room surfaces in an environmental chamber), and Martin and used in building thermal evaluations. This approach is similar to
262 S. Obyn, G. van Moeseke / Applied Thermal Engineering 87 (2015) 258e272

that performed in 2008 by Palyvos [48] who critically discusses the manual, i.e., 3.076 W/(m2 K) for internal surfaces [7] and 17.778 W/
available correlations in the literature for the external convection (m2 K) for external ones [14].
coefficients due to the wind and provides a suitable tabulation, on The geometry under consideration is a mechanically ventilated
the basis of algebraic form of the coefficients and their dependence office room of 3.6  5 m2 floor area and 2.7 m ceiling height (Fig. 3).
upon characteristic length and wind direction, in addition to wind Two different energy performance levels that are representative of
speed. The present work is also comparable to those of Khalifa in actual newly built offices in north and central Europe are consid-
2001 [36,37] and of Mirsadeghi et al. in 2013 [5] who reviewed the ered, as well as two internal gains levels (8.3 or 12.8 W/m2 in daily
literature respectively about internal and external CHTC values and average), two constructive modes (heavy and lightweight struc-
expressions. Finally, the present study also evaluate the assumption tures), and two room orientations (equator-facing or not). The en-
token by Ruivo et al. [46] and Yan et al. [47] about the consideration ergy performance levels differ according to the thermal resistance
of a constant convective heat transfer coefficient. of walls and windows (U ¼ 0.373 or 0.165 W/(m2 K) for external
walls and U ¼ 1.1 or 0.59 W/(m2 K) for windows). The surface
2. Method compositions are described in Appendix B. Internal and solar gains
are uniformly distributed in the room.
The aim of this study is to evaluate the impact of various The aim of studying several types of envelope, various orienta-
analytical and empirical expressions of CHTC when estimating tions and internal gains levels is to evaluate the he impact of CHTC
thermal loads of a building. Therefore, a thermal model of a stan- expression in a large range of cases in order to obtain general
dard office building is defined in TRNSYS 17 to test the impact of the conclusions and identify specific cases.
CHTC expressions on the annual heating and cooling demands and The climate file used represents an average climate year at Uccle
the maximal loads. Offices are often used as case studies to discuss (Brussels), Belgium, according to the Meteonorm database. This
CHTC values, such as in Awbi and Hatton (2000) [31] (experiments year includes 2172 heating degree-days and 85 cooling degree-days
conducted in a small-office-sized), Novoselac et al. (2006) [49] (calculated with 15  C and 18  C as reference values for heating and
(experiments in a full-scale facility representing an office space), cooling respectively). Temperatures vary between 7  C on winter
and Causone et al. (2009) [41] (experimental evaluation of heat nights and 30  C on the warmest day of summer. The daily tem-
transfer coefficients between radiant ceiling and room in a test perature shift is around 8  C. The cumulated daily global horizontal
chamber in typical conditions of occupancy of an office or resi- radiation varies between 50 and 7950 Wh/(m2 day).
dential building). Each expression of CHTC (corresponding to the Internal gain values used in the model are based on the works of
surface type) identified in the literature (presented in Appendix A) Massart and De Herde [50] (hypothesis of calculation for a tool to
is then applied to one type of surface (wall, window, ceiling, floor) help in the design of houses at very low energy consumption sus-
of a central office at a time. The value of h of other surfaces of the tainable new homes), Lie bard and De Herde [51] (data provided in
room equals constant default value provided in the TRNSYS manual their Treaty of bioclimatic architecture and urbanism), CIBSE guides
[7]. The aims are (1) to evaluate the range of value of CHTC obtained A and F [52,53], and the ASHRAE Handbook [54]. Two cases are
with identified expressions of CHTC for each type of surface and for distinguished: small and large internal gains. One person works in
the four categories presented in Section 1.1 and (2) to assess the the office for the small internal gains case and two people for the
impact of this variation on calculated annual and maximal heating large, with a typical work schedule in both cases. This produces
and cooling demands. mean values for internal gains in office rooms due to occupation
CHTC of internal surfaces of boundary rooms and of all external (occupants and equipment) of 16.4 W/m2 for small internal gains
faces are set at the constant default value provided in the TRNSYS and 29.1 W/m2 for large, and 6 W/m2 for lighting during occupation

Fig. 3. Geometry of the office room under consideration.


S. Obyn, G. van Moeseke / Applied Thermal Engineering 87 (2015) 258e272 263

(8:00 a.m. to 6:00 p.m. for lighting in office and corridors and 8:45 Table 3
a.m. to 5:15 p.m. for occupants and equipment in offices). Mean values of annual heating and cooling demands over the 1 year simulation
period in the case of standard insulation with heavy structure and large internal
Fresh air is mechanically supplied in the rooms and extracted in gains, with a southern orientation.
corridors. The operating airflow is 40 m3/(h pers) from 7:00 a.m. to
9:00 p.m. on workdays and 6 m3/(h pers) on nights and weekends. Surface Annual heating Annual cooling
demands [kWh/m2] demands [kWh/m2]
The ventilation installation is equipped with a heat recovery
system that pre-heats the pulsed air. A bypass system is placed Mean value Deviation Mean value Deviation

on the heat recovery system to perform free cooling when Wall Cat. 1 8.09 0.01 65.80 0.07
Tinside > Toutside > 16  C. Cat. 2 7.16 0.61 63.10 4.32
The office rooms are heated to 20  C from 6:30 a.m. to 8:00 p.m.; Cat. 3 5.70 1.08 50.55 7.15
Cat. 4 6.75 0.00 59.50 0.00
internal temperature may drop to 15.5  C nights and weekend. Note Floor Cat. 1 8.37 0.26 67.39 1.53
that the position of radiators in the offices is not defined in the Cat. 2 7.45 0.34 63.66 1.97
software. There is no heating element in corridors: they are heated Cat. 3 7.14 0.65 60.19 3.86
by the air transferred from the offices. The same schedule is Cat. 4 7.60 0.35 63.20 2.13
Ceiling Cat. 1 8.03 0.60 65.42 3.09
considered for cooling with a 25  C set point during weekdays and
Cat. 2 7.34 0.47 62.33 2.85
40  C for nights and weekends. There are no different heating Cat. 3 7.52 1.25 62.68 6.34
systems considered in this study. Only heating and cooling set Cat. 4 7.47 0.69 62.80 3.80
temperatures are considered in order to evaluate the impact of Window Cat. 1 8.08 0.00 65.75 0.01
Cat. 2 8.10 0.32 66.05 0.98
CHTC expression on energetic consumption evaluation of an office
Cat. 3 7.62 0.18 64.49 0.61
building. Cat. 4 7.90 0.00 65.56 0.00
Wall þ window Cat. 1 8.09 0.01 65.80 0.07
Cat. 2 6.92 0.91 61.96 5.86
Cat. 3 5.25 1.27 48.61 8.14
3. Results
Cat. 4 6.60 0.00 59.23 0.00

The first step consists in the establishment of the criteria to use


to compare the different values of CHTC provided by identified
All correlations of CHTC used in this study are presented in
correlations. The results are presented in two stages. The first
Appendix A. In each presentation of results, a brief explanation of
compares CHTC values obtained in the different cases studied,
the equations considered is given.
considering mean values and standard deviations. The second
The results presented in Tables 2e4 and Fig. 4 correspond to the
evaluates the impact of these values of CHTC on heating and cooling
case of standard insulation, southern orientation, heavy structure,
loads and annual demands. This evaluation is performed by setting
and large internal gains. This case study produces representative
all CHTC values as constant except for one type of surface (wall,
results and highlights each phenomenon observed. Although the
window, ceiling, floor) of the studied office. CHTC of internal sur-
results for other cases are not exhaustively presented, they are
faces of other offices and of all external faces are set at the constant
considered in the discussion.
default value provided in the TRNSYS manual, i.e., 3.076 W/(m2 K)
for internal surfaces [7] and 17.778 W/(m2 K) for external ones [14].
For example, the constant default value of the convective coeffi- 3.1. CHTC values
cient provided by the simulation software (TRNSYS) is set every-
where except on the ceiling of studied office, where all correlations Table 2 presents annual mean values of CHTC obtained with all
corresponding to this type of surface identified in the literature are expressions available in Appendix A, classified by type of surface
successively evaluated; i.e. in this case all expressions presented in and by category of correlation, in the case of standard insulation
Table A.3 in Appendix A. with heavy structure and large internal gains, with southern

Table 2
Annual mean values of CHTC obtained in the case of standard insulation with heavy structure and large internal gains, with a southern orientation. Mean values are calculated
for the 1 year simulation period and between correlations of the same category.

Surface considered Category of correlation Mean value [W/(m2 K)] Deviation [W/(m2 K)] Maximum value [W/(m2 K)] Minimum value [W/(m2 K)]

Wall Cat. 1 3.07 0.01 3.08 3.06


Cat. 2 1.72 0.56 2.63 0.40
Cat. 3 0.43 0.90 2.03 0.05
Cat. 4 5.87 0.00 5.87 5.87
Floor Cat. 1 3.90 0.78 4.60 3.06
Cat. 2 1.34 0.57 2.68 0.30
Cat. 3 0.90 1.52 3.94 0.00
Cat. 4 142.79 16.95 154.78 130.81
Ceiling Cat. 1 3.01 1.45 4.04 0.95
Cat. 2 1.21 0.96 2.78 0.28
Cat. 3 1.56 2.70 5.61 0.00
Cat. 4 166.53 184.14 364.35 0.09
Window Cat. 1 3.07 0.01 3.08 3.06
Cat. 2 3.25 2.37 8.25 0.53
Cat. 3 0.47 0.81 2.03 0.05
Cat. 4 3.26 0.00 3.26 3.26
Wall & window Cat. 1 3.07 0.01 3.08 3.06
Cat. 2 1.83 0.80 3.08 0.44
Cat. 3 0.43 0.90 2.03 0.05
Cat. 4 6.01 0.00 6.01 6.01
264 S. Obyn, G. van Moeseke / Applied Thermal Engineering 87 (2015) 258e272

Table 4
Mean and maximum values (per category of correlation) of maximum heating and cooling powers over the 1 year simulation period, in the case of standard insulation, heavy
structure, large internal gains, and southern orientation, for one surface at a time.

Surface Maximum heating powers [W/m2] Maximum cooling powers [W/m2]

Mean Deviation Maximum powers Mean Deviation Maximum powers

Wall Cat. 1 40.64 0.08 40.69 100.53 0.12 100.61


Cat. 2 35.36 4.05 41.60 100.46 8.28 112.08
Cat. 3 27.15 6.21 38.22 75.44 11.82 96.49
Cat. 4 31.49 0.00 31.49 93.83 0.00 93.83
Floor Cat. 1 42.32 1.53 43.48 103.48 2.84 106.06
Cat. 2 36.73 1.99 40.55 98.77 4.04 104.06
Cat. 3 35.23 3.81 42.83 91.18 6.40 103.82
Cat. 4 36.95 1.02 37.67 93.87 7.37 99.08
Ceiling Cat. 1 40.33 3.60 42.83 99.94 5.23 103.52
Cat. 2 36.05 2.74 40.33 95.70 5.24 104.00
Cat. 3 37.79 7.98 49.71 95.66 11.00 112.05
Cat. 4 44.16 12.87 58.02 94.39 5.90 230.98
Window Cat. 1 40.59 0.01 40.60 100.45 0.01 100.46
Cat. 2 40.89 1.57 43.86 101.86 2.07 105.54
Cat. 3 38.45 0.89 40.08 97.90 1.18 100.00
Cat. 4 39.92 0.00 39.92 100.56 0.00 100.56
Wall þ window Cat. 1 40.64 0.09 40.71 100.54 0.13 100.63
Cat. 2 34.18 5.68 42.77 99.48 11.56 114.65
Cat. 3 25.44 6.88 37.72 72.12 13.44 96.04
Cat. 4 35.03 0.00 35.03 93.97 0.00 93.97

orientation. The mean values presented in Table 2 correspond to the a visible impact on of heating and cooling demands, with
average of annual mean values of CHTC obtained with all expres- maximum deviations of 17.4% in mean values for heating and
sions of each category. Deviation values correspond to the deviation 12.5% for cooling. Mirsadeghi et al. [5] also observed deviations
between annual mean values of expressions of each category. in the values of yearly heating and cooling demand in function of
As expected, the results (Table 2) indicate that large differences the external CHTC expression considered. Note that, as explained
can appear between compiled correlations, as concluded by Khalifa in Section 2, the CHTC expressions are applied to one type of
in 2001 [36,37], and that some equations yield improbable results surface (wall, window, floor and ceiling) of the studied office at a
and illogical behavior; e.g., some expressions of category 4 (mixed time.
function) proposed by Fisher [26] and Beausoleil-Morrison [29,30]. Nevertheless, the very high values of CHTC observed by
This is a consequence of the formulation of these expressions as a considering expressions of category 4 for floor and ceiling don't
function of 1/DT. Therefore, when the temperature difference be- lead to significantly higher or lower values of annual heating and
tween surface and inside air is close to zero, the value of CHTC is cooling demand values than those calculated with other cate-
very high, which is unrepresentative of reality. This occurs mainly gories. This is due to the formulation of expressions of category 4
for horizontal surfaces (floor and ceiling). as a function of 1/DT which lead to punctually very high values of
On the other hand, expressions of hc as a function of airflow (cat. CHTC (till 100,000 W/(m2 K)). The high values of CHTC for ex-
3) give very different values of CHTC from other equations, except pressions of category 4 for floor and ceiling (see Table 2) are then
that of Fisher and Pedersen [27]. The minimum values obtained influenced by these occasional spikes. Moreover, these very high
with this category are much lower than those given by the corre- values of CHTC don't influence heat exchange between inside and
lation of constants or the function of the temperature difference. outside and then energy demands of the building, because when
Also, some correlations of category 2 (function of temperature they occur, the temperature difference between the air and the
difference) given by Khalifa [12], which are window specific, pro- surface is zero.
vide higher values of CHTC than those obtained with correlations of Although a detailed observation of the results shows a wider
vertical surfaces. This appears in the maximum values given in difference in terms of heating and cooling demands between
Table 2 (see also Fig. 1 in Section 1.1). correlation of category 3 (function of airflow) and others for
Finally, the results presented correspond to one particular case. “wall” and “wall þ window” surfaces, little difference is apparent
Note, however, that the conclusions are the same for each case if windows surfaces are considered alone, probably because of the
studied (i.e., for alternative constructive mode, orientation, ener- smaller surface concerned. For floor and ceiling, category 3 ex-
getic performance, or internal gains level). pressions give results which are closer to those of other cate-
gories. Indeed, annual mean values of CHTC obtained with this
3.2. Impact on annual heating and cooling demands category of expression for such surfaces show less variation from
other categories than do those for other surfaces, as shown in
Table 3 presents the values for heating and cooling demands Table 2.
obtained by considering each expression available in Appendix A. By only considering correlations of categories 1 (constant value)
This way, annual demands presented in Table 3 correspond to the and 2 (function of temperature difference), despite significant de-
average of annual demands obtained considering all expressions viations of CHTC values (see Table 2) e with a maximum value of
of each category. Deviation values correspond to the deviation 79.3% for expression of category 2 applied to the ceiling e standard
between annual demand values of each category. Mean values deviations of heating and cooling demands are limited, with
are calculated by type of surface and category of correlation, maximum values of 13.2% for heating demand and 9.5% for cooling
given standard insulation, heavy structure, large internal gains, demand, in the case of expression of category 2 applied to
and southern orientation. It shows that expressions of CHTC have “wall þ window”.
S. Obyn, G. van Moeseke / Applied Thermal Engineering 87 (2015) 258e272 265

Fig. 4. Up e Daily mean heating (>0) and cooling (<0) loads; Down e Distribution of temperature difference between the inside air and surface studied during the year for
representative expressions of CHTC in the case of standard insulation, heavy structure, large internal gains, and southern orientation: left e on floor and right e on ceiling.

When considering all simulated cases, not only the one detailed the maximum deviation calculated is of 9.1% in heating demands in
in Table 3, the maximum values of standard deviation calculated are the case of a high-insulation heavy building with small internal
18.3% for annual heating demands and 10.9% for annual cooling gains and southern orientation and of 5.9% in cooling demands in
demands in the case of a standard-insulation heavy building with the case of a standard-insulation heavy building with large internal
small internal gains (northern orientation for heating and southern gains and northern orientation. Also, values of deviation greater
orientation for cooling demands), which is significant but far less than 8% appear in only one out of four cases for heating demands. In
than the maximum value of deviation of CHTC, which is 80.9% for other simulated cases, the deviation is mainly lower than 5%. It
the expression of category 2 applied to the ceiling in the case of a seems therefore that a constant value for CHTC can reasonably be
high-insulation heavy building with small internal gains and used when the desired result is an annual report. But does the
southern orientation. TRNSYS value yield the most representative results? In fact, the
Based on these results, the relevance of a constant CHTC can be value of CHTC given in the TRNSYS manual [7] is greater than the
evaluated, as it was the assumption of several works as those of average observed CHTC. However, other constant values of
Ruivo et al. in 2013 [46] and of Yan et al. in 2015 [47]. The convective coefficient available in the literature (cat. 1.1) are of the
confrontation of values corresponding to the constant value given in same order or higher.
the TRNSYS manual [7], i.e., 3.076 W/(m2 K), and the mean values of
annual energetic demands obtained with correlations of categories 3.3. Impact on heating and cooling loads
1 and 2 (constant values and correlation functions of DT), in the case
presented in Table 3, yields a deviation of 7.1% in heating demands Fig. 4 shows heating and cooling mean loads observed in the
and 2.9% in cooling demands. Considering other simulation cases, office room in function of temperature difference between ambient
266 S. Obyn, G. van Moeseke / Applied Thermal Engineering 87 (2015) 258e272

air and the surface studied for representative expressions of cate- It also appears that, as with annual energy demands, correla-
gories 1 (TRNSYS manual [7]), 2 (Rogers and Mayhew & ASHRAE tions of category 3 (function of airflow) applied to walls and
[17,18]), 3 (Fisher [26,27] (a)), and 4 (Beausoleil-Morrison [29,30] walls þ windows provide mean values for heating and cooling
(a)) for floors and ceilings in the case of standard insulation, loads that significantly differ from other categories of expressions.
heavy structure, large internal gains, and southern orientation. Finally, if studying the relevance of a constant CHTC, as con-
In Fig. 4 (a), lines up to 0 represent the trend for daily mean siderd by Ruivo et al. in 2013 [46] and by Yan et al. in 2015 [47], i.e.,
values of heating loads and the lines under 0 cooling loads, in 3.076 W/(m2 K) as proposed in the TRNSYS manual [7], by
function of temperature difference between inside air and the comparing the results with mean values of annual maximum load
surface studied, for the representative expression of each category. obtained with correlations of categories 1 and 2 (constant values
Note that the values of the coefficient of determination R2 go as and correlations function of DT), in the case presented in Table 4,
high as 0.77 for heating curves and 0.97 for cooling curves for floor the values deviate 7.8% for heating loads and 2.6% for cooling loads.
surfaces and to 0.74 and 0.96, respectively, for ceiling surfaces. Considering other simulation cases, the maximum deviation
In Fig. 4 (a) left, it appears that the difference in heating and calculated is 9.8% in heating loads in the case of high-insulation,
cooling loads between the four categories of CHTC correlations heavy structure, small internal gains, and southern orientation
applied to the floor surface is significant, as observed by Mirsa- and 3.3% in cooling loads in the case of standard insulation, heavy
deghi et al. in 2013 for external CHTC [5]. However, correlations structure, small internal gains, and northern orientation. Further-
from the TRNSYS manual [7] (cat. 1) and Beausoleil-Morrison more, the deviation value is always less than 2% in the case of light
[29,30] (a) (cat. 4) provide very close results in terms of daily structure and less than 4% in 63% of cases for heavy structure. The
mean values of cooling powers in function of temperature differ- reference value of CHTC may then reasonably be used to evaluate
ence between inside air and the floor of the room. On the other heating and cooling loads.
hand, the correlations from Fisher [26] (a) (cat. 3) generally lead to
smaller values for mean heating and cooling powers (see Fig. 4 (a) 4. Conclusions and outlook
left and Table 4).
In the case of the ceiling surface presented in Fig. 4 (a) right, To evaluate the impact of the expression of CHTC on heating and
categories 2 (Rogers and Mayhew & ASHRAE [17,18]) and 3 (Fisher cooling demands, a thermal model of a standard office building is
[26,27] (a)), expressions provide very close values of daily mean defined in TRNSYS 17. The criteria considered in this study are the
power in function of temperature difference between inside air and mean value of CHTC over the year, heating and cooling annual
ceiling. The same holds true for categories 1 (TRNSYS manual [7]) demands and loads, and deviation of CHTC and annual demands.
and 4 (Beausoleil-Morrison [29,30]). These trends do not, however, Considering one type surface at a time (wall, window, floor,
appear in values of annual heating and cooling demands (presented ceiling), it firstly appears that large differences occur in the annual
in Tables 3 and 4) because of the influence of CHTC correlation on mean value of CHTC between CHTC expressions considered, as
surface temperature. Fig. 4 (b) presents the distribution of daily observed by Khalifa in 2001 [36,37]. Secondly, expressions of CHTC
temperature differences between inside air and floor (left) or ceil- expressed as a function of airflow (cat. 3) give values of annual
ing (right). heating and cooling demands that diverge from those provided by
Finally, Table 4 shows annual mean and maximum values per other categories of expressions when applied to vertical surfaces.
type of surface for maximum calculated heating and cooling loads Also, mixed functions of DT may lead to illogical behavior due to
(see Appendix A) for standard insulation, heavy structure, large their formulation as a function of 1/DT. Therefore, when the tem-
internal gains, and southern orientation. Mean maximum loads perature difference between surface and inside air is close to zero,
values presented in Table 4 correspond to the average of maximum the value of CHTC is very high, which is not representative of reality.
loads obtained considering all expressions of each category. Devi- This applies mainly to horizontal surfaces (floor and ceiling).
ation values correspond to the deviation between maximum load Regarding mean values of annual heating and cooling demands,
values obtained in each category. It shows substantial differences calculated with either a constant CHTC value or correlations as a
by changing the category of CHTC expression in terms of maximum function of temperature difference between surfaces and air, it
heating (up to 30.5%) and cooling (up to 55.2%) powers (see appears that, despite significant deviations in CHTC values
Table 4). Otherwise, the maximum deviation observed in mean (maximum deviation of 80.9% for expressions of category 2 applied
maximum loads per category of expression of CHTC and per surface to ceiling surface in the case of high-insulation, heavy structure,
is 29.2% for heating loads and 56.4% for cooling loads. All simula- small internal gains, and southern orientation), standard deviations
tions considered, it appears that the range of these deviations (of heating and cooling demands) are quite limited (maximum
varies with structure type for heating loads and with orientation for values of 18.3% for annual heating demands and 10.9% for annual
cooling loads. This maximum deviation is always lower than 10% cooling demands in the case of standard insulation, heavy struc-
for north-oriented office cooling loads and for light structure ture, and small internal gains) but not insignificant. This coefficient
heating loads. Otherwise, the maximum deviations observed are then has a real impact on the evaluation of the energetic perfor-
45.8% for heating loads in the case of high-insulation, heavy mance of an office building, which is a similar conclusion to those of
structure, large internal gains, and northern orientation, and 56.4% Mirsadeghi et al. for external CHTC [5].
for standard insulation, heavy structure, large internal gains, and Moreover, all simulation cases considered, there is a maximum
southern orientation. For both orientations, the range of value de- deviation of annual energy demands of 9.1% between (1) the
viations varies in function of structure type. Deviations in the mean default constant CHTC value from the TRNSYS manual [7] and (2)
maximum heating and cooling load per category of expression and the mean values obtained with expressions of categories 1 (hc ¼ cst)
type of surface are respectively smaller than 6.6% and 13.3% for light and 2 (hc ¼ f(DT)). This maximum deviation occurs for heating
structures, but they rise to 45.8% and 56.4% for heavy structures. For demands in the case of high insulation, heavy structure, small in-
heavy structures, the expression of CHTC has a major impact on ternal gains, and southern orientation. The maximum difference in
HVAC installation design and cost. cooling demands between results obtained with the TRNSYS
S. Obyn, G. van Moeseke / Applied Thermal Engineering 87 (2015) 258e272 267

expression and mean results with all expressions of types 1 and 2 is -


5.9% and occurs in the case of lower insulation, heavy structure, large deviations appear in the evaluation of maximum heating
large internal gains, and northern orientation. and cooling powers only in the case of offices with heavy struc-
Moreover, differences between constant expressions of CHTC ture (deviation < 13.3% for light structure cases) but have a major
and other expressions that are greater than 8% only appear in 25% of influence on design decisions and the cost of HVAC systems in
cases of heating demands. In other cases, the deviation is generally buildings;
lower than 5%. The benefit of using sophisticated expressions of - if the constant value of CHTC from the TRNSYS manual [7] may
CHTC (cat. 2, 3, or 4) seems limited except in the case of heavy reasonably be used to make an annual report of heating and
structure and small internal gains. cooling demands, it may also be considered to provide an
Finally, if the correlations of CHTC seem to have an important overview of annual maximum heating and cooling loads in the
impact on values of annual heating and cooling demands, they also case of the office building considered in this study;
affect values of maximum heating and cooling powers. A maximum - if category 1 (hc ¼ constant) and category 2 (hc ¼ f(DT)) ex-
variation of mean maximum power per type of expression of CHTC pressions show little divergence, categories 3 (hc ¼ f(airflow))
and per type of surface equivalent to 45.8% in heating power and and 4 (hc ¼ mixed function of DT) lead to significant differ-
56.4% in cooling power is observed. Large deviations appear only ences in energy loads. It seems therefore recommendable to
for offices with heavy structure (deviation < 13.3% for light struc- determine if convection will be buoyant or forced before
ture cases) but have a major influence on design decisions and the selecting an expression of CHTC for a building's thermal
cost of HVAC systems in buildings. simulation.
Considering the relevance of using the reference constant value
of CHTC available in TRNSYS [7], a maximum deviation of maximum It is to be noted that in this study solar gains have not been
heating load of 9.8% is observed between (1) the constant value of localized. They are calculated as uniformly distributed on all inte-
CHTC from TRNSYS manual [7] and (2) the mean values obtained rior surfaces of the area studied. In a later study, the impact of
with expressions of categiries 1 (hc ¼ constant) and 2 (hc ¼ f(DT)). localized solar gains on CHTC values and then on annual heating
For cooling power, the maximum deviation is 3.3%. Therefore, if the and cooling demands can be measured. Furthermore, the present
constant value of CHTC from the TRNSYS manual [7] may reason- global evaluation excludes correlations of CHTC function with
ably be used to make an annual report of heating and cooling de- nozzle position and air velocity at nozzle opening. In a more
mands, it may also be considered to provide an overview of annual detailed study, it would be possible to localize and measure the
maximum heating and cooling loads in the case of the office ventilation nozzle and integrate these correlations within the
building considered in this study. study. Finally, the present study is theoretical and evaluates the
In any event, if category 1 (hc ¼ constant) and category 2 impact of CHTC correlation on one type of surface at a time. No
(hc ¼ f(DT)) expressions show little divergence, categories 3 measurements have been carried out on a real test building. In situ
(hc ¼ f(airflow)) and 4 (hc ¼ mixed function of DT) lead to significant measurements would permit to define the most appropriate
differences in energy loads. It seems therefore recommendable to expression to use in simulation for several cases. The study high-
determine if convection will be buoyant or forced before selecting an lights the most sensitive case for which a large range of energetic
expression of CHTC for a building's thermal simulation. Also, ex- demands is calculated and then for which an appropriate CHTC
pressions of category 3 are probably better adapted to the case of should be defined. A study considering all surfaces simultaneously
night cooling with a high level of air change per hour. Expressions of would provide more practical results with greater relevance for the
category 4 may be used if the difference of temperature between construction sector.
inside air and surface is sufficient, but they may not be appropriate
for building simulation due to their formulation as a function of 1/DT. Acknowledgements
Summarizing, the main conclusions of this study are:
This research was conducted in the context of the SIMBA project
- large differences occur in the annual mean value of CHTC be- funded by the European FEDER program.
tween CHTC expressions considered, as observed by Khalifa in
2001 [36,37];
- mixed functions (category 4) lead to illogical behavior due to Appendices
their formulation as a function of 1/DT;
- the CHTC has a real impact on the evaluation of the energetic A. CHTC correlations studied
performance of an office building, as concluded by Mirsadeghi
et al. for external CHTC [5]; Stable and unstable refers to the direction of heat flow and the
- the benefit of using sophisticated expressions of CHTC (cat. 2, 3, associated buoyancy relative to the surfaces. Stable is when the
or 4) in place of a constant value of CHTC, as considered by Ruivo natural tendency is to retard flow in the sense that rising warmer
et al. [46] and Yan et al. [47], for the evaluation of annual cooling air, or falling cooler air, is driven against the surface. Unstable is
and heating energy demands seems limited except in the case of when the natural tendency is to enhance flow in the sense that
heavy structure and small internal gains; rising warmer air, or falling cooler air, is free to move away from the
surface.
268
Table A.1
Vertical surface e Natural convection (NC); Forced convection (FC), and Mixed convection (MC).

Reference Correlations [W/(m2 K)] Remarks

ASHRAE [6] hc ¼ 3:076 NC


TRNSYS manual [7] hc ¼ 3:055556 NC and FC, Internal surfaces
Awbi and Hatton [9] (a) hc ¼ 1:823
Dh0:121
$ðDTÞ0:293 NC, Experimental study on actively heated walls, Test chamber: 2.78  2.78  2.30 m3
Khalifa and Marshall [10] hc ¼ 1:983$ðTÞ1=4 Experimental study on interior building surfaces: Cold vertical wall in a test cell
(2.35  2.95  2.08 m3) heated by a small fan heater (the wall not in front of the
fan heater), DT  5  C, Fan power ¼ 20 W
Khalifa & Khalifa and (a) hc ¼ 2:35$ðDTÞ0:21 Experimental study on interior building surfaces using a real-sized indoor test cell
Marshall [11] (b) hc ¼ 2:20$ðDTÞ0:22 (2.35  2.95  2.08 m3). Valid for a room with (a) a radiator under glazing and
DT  5  C, (b) a radiator opposite to the cold wall and DT  5  C
Khalifa & Khalifa and (a) hc ¼ 2:07$ðDTÞ0:23 Experimental study on interior building surfaces using a real-sized indoor test cell
Marshall [11,12] (b) hc ¼ 2:30$ðDTÞ0:24 (2.35  2.95  2.08 m3)

S. Obyn, G. van Moeseke / Applied Thermal Engineering 87 (2015) 258e272


(a) NC and FC. Valid for a room heated by a circulating fan heater (for surfaces not
opposite fan), with a heated floor, with a heated edge or heated by a radiator
(not under a window); for surfaces of walls away from the heat source and
DT  5  C.
(b) NC. Valid for a room with a heated vertical wall(s) (not applicable for the heated
walls) and a room heated by a radiator under a window and DT  5  C
Khalifa [12] (a) hc ¼ 1:98$ðDTÞ0:32 Experimental study on interior building surfaces using a real-sized indoor test cell
(b) hc ¼ 2:92$ðDTÞ0:25 (2.35  2.95  2.08 m3). NC.
Valid (a) for a room heated by radiator (not under a window); for inside surfaces of
walls near the heat source (adjacent to radiator) and (b) for a room heated by
circulating fan heater (only for surfaces opposite fan)
Min et al. [13] (a) hc ¼ 1:517$ðDTÞ1=3 Quote from Ref. [55]. Correlations converted from British units to SI units. MC
(b) hc ¼ 2$DT
0:32

H0:04
(Test cell: 3.6  7.2 or 3.6  3.6 m2). For floor heated case
TRNSYS manual [14] hc ¼ 0:41667$ðDTÞ1=4 MC
Hatton and Awbi [15] hc ¼ 1:57$ðDTÞ0:31 MC (Test cell: 2.78  2.78  2.3 m3)
Alamdari and Hammond [16] (a) hn  1=4 o i1=6 Natural convection and radiation in a panel-heated room
6
hc ¼ 1:50$ DT H þ f1:23$ðDTÞ1=3 g6 Correlated data from other investigators. Valid for DT  600  C
Rogers and Mayhew [17]* hc ¼ 1:42$ðDTÞ1=4 Laminar NC
& ASHRAE [18]** & ISO 15099 hc ¼ 1:31$ðDTÞ1=3 Turbulent * Quote from Refs. [56]; ** Correlations converted from British units to SI units
1=4
CIBSE Guide C3 [19] hc ¼ k$0:48
H $ðGrÞ Laminar e
1=3
hc ¼ k$0:119
H $ðGrÞ Turbulent
1=4
Li et al. [22] (a) hc ¼ 3:08$ðDTÞ Experimental study (Room 3.40  4.00  2.60 m3).
(b) hc ¼ 2:88$ðDTÞ 1=4
Valid (a) for DT  1.5  C in an occupied office room under normal working conditions.
Include the effect of some thermal radiation and ventilation
Valid (b) for DT  1.5  C in an occupied office room under normal working conditions.
Include the effect of ventilation
Fohanno and Polidori [23]  1=4 NC. Internal vertical building surface uniformly heated. Expressions derived from a
hc ¼ 1:332$ DT
H for Ra*H  6; 3$109 theoretical analysis based on the integral formalism for simple buoyancy flow
conditions
hc ¼ 1:235$eð0:0467 HÞ $½DT0:316 for Ra*H  6; 3$109
hc ¼ k $ð24:8 þ 0:36 Re0:8
Fisher [26] 1=3
Vroom e Þ MC. Experimental investigation of mixed convection heat transfer in a rectangular
enclosure. Valid for 10  C  Tdiffuser  25  C and 3  ACH  100
Fisher [26,27] (a) hc ¼ 0:190$ðACHÞ0:8 MC. (a) Experimental study and (b) Ceiling jets in isothermal rooms, (c) Free
(b) hc ¼ 0:199 þ 0:190$ðACHÞ0:8 horizontal jets in isothermal rooms
(c) hc ¼ 0:110 þ 0:132$ðACHÞ0:8
Fisher and Pedersen [27] hc ¼ 1:208 þ 1:012$ðACHÞ0:604 FC with ceiling diffuser. Equation from laboratory chamber measurements
Beausoleil-Morrison [29,30] (a) nh  1 i6 h i oð3=6Þ 
3 1=3 MC. Internal building surface
1 6
4
Tsurf  Tdiffuser
hc ¼ 1:5$ DT
H þ 1:23$DT 3 þ DT $½0:199 þ 0:190$ðACHÞ0:8  If flow driving forces from mechanical forces are augmented by the driving forces
by buoyancy
Combined correlations originally developed by Refs. [16,27]
Table A.2
Horizontal surface FLOOR e Natural convection (NC), Forced convection (FC), and Mixed convection (MC).

Reference Correlations [W/(m2 K)] Remarks

ASHRAE [6] hc ¼ 0:948 For reduced convection NC


hc ¼ 4:040 For enhanced convection
TRNSYS manual [7] hc ¼ 3:055556 Constant value on internal surfaces
Chen et al. [8] hc ¼ 4:7 For DT > 0 K MC. Ventilated room (Test chamber: 5.6  3.0  3.2 m3) with heated surfaces with
hc ¼ 4:3 For DT < 0 K stably stratified conditions mixed convection (laminar and turbulent flows) for a relatively low Reynolds number
Awbi and Hatton [9] (a) hc ¼ 2:175
D0:076
$ðDTÞ0:308 NC. Experimental study (Test cell: 2.78  2.78  2.30 m3). Average CHTC for an
h
actively heated floor in a room, for 9  108 < Gr < 7  1010
Khalifa and Marshall [11] (a) hc ¼ 3:10$ðDTÞ0:17 Experimental study on interior building surfaces using a real-sized indoor test cell
(b) hc ¼ 2:27$ðDTÞ0:24 (2.35  2.95  2.08 m3). Valid for DT  5  C
Min et al. [13] (a) hc ¼ 1:931$ðDTÞ1=3 Quote from Ref. [55]. Converted from British units to SI units
TRNSYS manual [14] hc ¼ 0:58611$ðTsurf  Tair Þ0:31 for Tsurf > Tair MC
hc ¼ 0:51944$ðTsurf  Tair Þ0:25 for Tsurf < Tair

S. Obyn, G. van Moeseke / Applied Thermal Engineering 87 (2015) 258e272


Alamdari and Hammond [16]   1=4 6
1=6 Buoyant flow, based on the data of [25,57,58] using [32] mathematical model. (a) NC
(a) hc ¼ 1:4$ DT
Dh þ ½1:63$ðDTÞ1=3 6 valid for 104  Ra  1012 and Tsurface > Tair and (b) NC and FC for stably stratified
!1=5 horizontal surface. Valid for Tair > Tsurface
(b) hc ¼ 0:6$ DT
D2h

Rogers and Mayhew [17]*  1=4 * Quote from Ref. [56]


& ASHRAE [18]** hc ¼ 1:32$ DT
L Laminar ** Correlation converted from British units to SI units
hc ¼ 1:52$ðDTÞ1=4 Turbulent
CIBSE Guide C3 [19]  1=4 e
hc ¼ 1:40$ DT
L Laminar

hc ¼ 1:70$ðDTÞ1=3 Turbulent
 
Schlapmann [21] Experimental study (Test cell: 4.00  4.95  2.70 m3)
hc ¼ 0:33$HL þ 0:75 $ðDTÞ1=3
A floor heated by warm water, vertical walls at different temperatures. Converted
from British to SI units
Walton [24] hc ¼ 9:482$ðDTÞ
1=3
for Tsurf < Tair NC and FC. Equation obtained by fitting curves from various sources
6:283
1=3
hc ¼ 1:810$ðDTÞ
2:382 for Tsurf > Tair
Yousef et al. [25] (a) hc ¼ k$0:622 $ðRaÞ 1=4
for 3$106  Ra  4$107 Upward-facing isothermal horizontal surface (a) average, (b) quote from Ref. [59],
L
hc ¼ k $0:162 $ðRaÞ 1=3
for Ra > 4$107 (c) quote from Ref. [58] and (d) quote from Ref. [60]
L
(b) hc ¼ k $ L0:71$ðRaÞ1=4 Laminar
hc ¼ k $0:17
L $ðRaÞ
1=3
Turbulent
1=4
(c) hc ¼ k$ 0:70L $ðRaÞ for 2$105  Ra  4$107
hc ¼ k $ 0:155 $ðRaÞ 1=3
for 4$107  Ra  109
L
(d) hc ¼ k $0:20L $ðRaÞ 1=3
for Ra > 2$105
Fisher [26] hc ¼ 1k $ð24:8 þ 0:36 $Re0:8 e Þ
Experimental investigation of mixed convection heat transfer in a rectangular
3
Vroom enclosure. Valid for 10  C  Tdiffuser  25  C and 3  ACH  100
Fisher [26,27] (a) hc ¼ 0:13$ðACHÞ0:8 MC with ceiling jets in isothermal rooms experimental study
Fisher [26,27] (b) hc ¼ 0:159 þ 0:116$ðACHÞ0:8 MC with ceiling jets in isothermal rooms
Fisher [26,27] (c) hc ¼ 0:704 þ 0:168$ðACHÞ0:8 MC with free horizontal jets in isothermal rooms
Fisher and Pedersen [27] hc ¼ 3:873 þ 0:082$ðACHÞ0:98 FC with ceiling diffuser
!0:8 Equation from laboratory chamber measurements
_
Goldstein and Novoselac [28] hc ¼ 0:048$ VL FC for perimeter zones with highly glazed spaces served by overhead
slot-diffuser-based air systems
Equation from laboratory chamber measurements
Beausoleil-Morrison [29,30]   1 6 h i
ð3=6Þ 
3 1=3 Mixed convection with the flow driving forces include both mechanical forces and
1 6
4
Tsurf  Tdiffuser
(a) hc ¼ 1:4$ DT
Dh þ 1:63$DT 3 þ DT $½0:159 þ 0:116 $ðACHÞ0:8  thermally unstable buoyancy
( !1 )3 
3 !1=3 Combined correlations originally developed by Alamdari and Hammond [16] and
5
DT Tsurf  Tdiffuser Fisher and Perdersen [27]
(b) hc ¼ 0:6$ D2h
þ DT $½0:159 þ 0:116$ðACHÞ0:8 

269
270
Table A.3
Horizontal surface CEILING e Natural convection (NC), Forced convection (FC), and Mixed convection (MC).

Reference Correlations [W/(m2 K)] Remarks

ASHRAE [6] hc ¼ 0:948 for reduced convection NC


hc ¼ 4:040 for enhanced convection
TRNSYS manual [7] hc ¼ 3:055556 Constant value
Chen et al. [8] hc ¼ 4:0 Ventilated room (5.6  3.0  3.2 m3) with heated surfaces with mixed
convection (laminar and turbulent flows) for a relatively low Reynolds
number and DT < 0 K
Awbi and Hatton [9] (a) hc ¼ 0:704 $ðDTÞ0:133 Experimental study (Test chamber: 2.78  2.78  2.30 m3)
D0:601
h
(a) Heated ceiling in a chamber for 9.108  Gr  1011
(b) hc ¼ 1:736
D0:52
$ðDTÞ0:16
h (b) Average CHTC of small heated plates on a chamber ceiling
(characteristic length determined by heated surface area only),
for 9.108  Gr  1011
Khalifa and Marshall [11] (a) hc ¼ 2:72$ðDTÞ0:13 Experimental study on interior building surfaces using a real-sized indoor

S. Obyn, G. van Moeseke / Applied Thermal Engineering 87 (2015) 258e272


(b) hc ¼ 2:27$ðDTÞ0:24 test cell (2.35  2.95  2.08 m3). NC and FC for DT  5  C. Valid for a room
convectively heated by a circulating fan heater (for surfaces not
opposite fan), with a heated floor or edge or heated by a radiator
(not under a window); for inside surfaces away from the heat source
Khalifa & Khalifa and Marshall [11,12] hc ¼ 3:10$ðDTÞ0:17 Experimental study on interior building surfaces using a real-sized indoor
test cell (2.35  2.95  2.08 m3). Valid for a room with a heated vertical
wall(s) and a room heated by a radiator under a window. NC for DT  5  C
Min et al. [13] (a) hc ¼ 1:034 $ðDTÞ1=3 Quote from Ref. [55]
Correlation converted from British units to SI units
TRNSYS manual [14] hc ¼ 0:58611 $ðTsurf  Tair Þ0:31 for Tsurf > Tair MC
hc ¼ 0:51944 $ðTsurf  Tair Þ0:25 for Tsurf < Tair
Alamdari and Hammond [16]   1=4 6
1=6 NC (a): For floor surfaces in a buoyant thermal situation. Based on the
(a) hc ¼ 1:4 $ DT
Dh þ ½1:63 $ðDTÞ1=3 6 data of [57], [58] and [25], using [32] mathematical model. Valid for
!1=5 Tsurf < Tair and 104  Ra  1012. NC and FC (b) for stably stratified
(b) hc ¼ 0:6$ DT horizontal surface and floor surfaces in a buoyant thermal situation. Based
D2h
on the data of [61], [57], [62] and [63]. Valid for Tsurf > Tair and
104  Ra  1012
Min et al. [13]* & Rogers and  1=4 * Quote from Refs. [64], for a 0.60 m square plate
Mayhew [17]** & ASHRAE [18]*** hc ¼ 0:586 $ DT
L for DT  555  C ** Quote from Ref. [56]
*** Correlation converted from British units to SI units
CIBSE Guide C3 [19]  1=4 e
hc ¼ 0:64$ DT
L

Karadag [20] hc ¼ 3:1$ ðDTÞ0:22 NC. Correlation from numerical methods for chilled-ceiling
Walton [24] (a) hc ¼ 9:482$ðDTÞ
1=3
NC and FC. Correlation obtained by fitting curves from various sources.
6:283
(a) for Tsurf > Tair on unstable horizontal planes, (b) for Tsurf < Tair on
1:810 $ðDTÞ1=3
(b) hc ¼ 2:382 stable horizontal planes
Fisher [26] hc ¼ k
1=3 $ð24:8 þ 0:36$Re0:8
e Þ
Experimental investigation of mixed convection heat transfer in a
Vroom
rectangular enclosure. Valid for 10  C  Tdiffuser  25  C and 3  ACH  100
Fisher [26,27] (a) hc ¼ 0:49 $ðACHÞ0:8 (a) Experimental study (b) for MC with ceiling jets in isothermal rooms,
(b) hc ¼ 0:166 þ 0:484$ðACHÞ0:8 (c) for MC with free horizontal jets in isothermal rooms
2:8
(c) hc ¼ 0:064 þ 0:004 44$ðACHÞ
DTvent
Fisher and Pedersen [27] hc ¼ 2:234 þ 4:099 $ðACHÞ 0:503 FC with ceiling diffuser
Equation from laboratory chamber measurements
Beausoleil-Morrison [29,30] (a)   1=4 6
1=6 
3 1=3
Tsurf  Tdiffuser
hc ¼ 1:4 $ DT
Dh þ ½1:63 $ðDTÞ1=3 6 þ DT $½0:166 þ 0:484 $ðACHÞ0:8 

Beausoleil-Morrison [29,30] (b) ( ! 1 )3 


3 !1=3 Mixed convection with the flow driving forces include both mechanical
5
DT Tsurf  Tdiffuser
hc ¼ 0:6$ D2h
þ DT $½0:166 þ 0:484 $ðACHÞ0:8  forces and thermally stable buoyancy
Combined correlations originally developed by Refs. [16] and [27]
S. Obyn, G. van Moeseke / Applied Thermal Engineering 87 (2015) 258e272 271

Table A.4
Window e Natural convection (NC), Forced convection (FC) and Mixed convection (MC).

Reference Correlations [W/(m2 K)] Remarks

[6,7,9,13e19,22,23,26,29,30] Same equation as for vertical surface Other correlations are specific to walls
Khalifa [12] (a) hc ¼ 7:61 $ðDTÞ0:06 In room heated by radiator not located under the window
Experimental study
Khalifa [12] (b) hc ¼ 8:07 $ðDTÞ0:11 In room heated by radiator located under the window
Experimental study
Goldstein and Novoselac [28] !0:8 FC and MC
V_ diffuser
hc ¼ 0:117 $ length Correlation from laboratory chamber measurements
For zones with highly glazes surfaces and served by overhead
For WWR < 50% with window in upper part of wall slot-diffuser-based air system
!0:8
V_ diffuser
hc ¼ 0:093 $ length

For WWR < 50% with window in lower part of wall


!0:8
V_ diffuser
hc ¼ 0:103 $ length

For WWR > 50%

Table B.1
Surface compositions.

Insulation_structure Standard_lightweight Standard_heavy High_lightweight High_heavy

Material Thickness Material Thickness Material Thickness Material Thickness


[mm] [mm] [mm] [mm]

External roof Interior finishing 22 Concrete 300 Interior finishing 22 Concrete 300
Air layer / Insulation wool 110 Air layer / Insulation wool 245
Concrete 300 Waterproof membrane 5 Concrete 300 Waterproof membrane 5
Insulation wool 80 Insulation wool 220
Waterproof membrane 5 Waterproof membrane 5
U [W/(m2 K)] 0.301 0.284 0.137 0.136
Adjacent floor/ceiling Interior finishing 22 Concrete 300 Interior finishing 22 Concrete 300
Air layer / Air layer / Air layer / Air layer /
Concrete 300 Concrete 200 Concrete 300 Concrete 200
Air layer / Air layer /
Interior finishing 22 Interior finishing 22
Ground floor Interior finishing 22 Concrete screed 100 Interior finishing 22 Concrete screed 50
Air layer / Concrete 300 Air layer / Concrete 300
Concrete screed 50 Insulation wool 50 Concrete screed 50 Insulation wool 100
Concrete 300 Concrete 300
Insulation wool 30 Insulation wool 80
2
U [W/(m K)] 0.531 0.547 0.302 0.307
External wall Interior finishing 22 Concrete 250 Interior finishing 22 Concrete 250
Air layer / Insulation wool 80 Air layer / Insulation wool 200
Concrete 250 Waterproof membrane 5 Concrete 250 Waterproof membrane 5
Insulation wool 60 Insulation wool 180
Waterproof membrane 5 Waterproof membrane 5
U [W/(m2 K)] 0.373 0.380 0.164 0.165
Adjacent wall Interior finishing 22 Concrete 200 Interior finishing 22 Concrete 200
Air layer / Air layer /
Interior finishing 22 Interior finishing 22
Window double glazing double glazing triple glazing triple glazing
U [W/(m2 K)] 1.100 1.100 0.590 0.590

B. Surface compositions

Table B.2
Material descriptions.

Material Conductivity Capacity Density


References
[W/(m K)] [kJ/(kg K)] [kg/m3]

Concrete 22.68 1 2300 [1] J.L.M. Hensen, R. Lamberts, C.O.R. Negrao, A view of energy and building
Insulation wool 0.454 1 20 performance simulation at the start of the third millennium, Energy Build. 34
Waterproof membrane 1.80 1.4 1000 (2002) 853e855.
Interior finishing 0.576 1 200 [2] J.L.M. Hensen, R. Lamberts, Building Performance Simulation for Design and
Concrete screed 22.68 1 2300 Operation, Routledge, London, UK, 2011.
[3] A.M. Malkawi, G. Augenbroe, Advanced Building Simulation, Spon Press, New
Air layer Resistance ¼ 0.0125 m2 K/W
York, USA, 2003.
272 S. Obyn, G. van Moeseke / Applied Thermal Engineering 87 (2015) 258e272

[4] W.C. Turner, S. Doty, Energy Management Handbook, seventh ed., The Fair- [35] I. Beausoleil-Morrison, The adaptive simulation of convective heat transfer at
mont Press, Lilburn, USA, 2009. internal building surfaces, Build. Environ. 37 (2002) 791e806.
[5] M. Mirsadeghi, D. Co  stola, B. Blocken, J.L.M. Hensen, Review of external [36] A.-J.N. Khalifa, Natural convective heat transfer coefficient e a review: I.
convective heat transfer coefficient models in building energy simulation Isolated vertical and horizontal surfaces, Energy Convers. Manage. 42 (2001)
programs: implementation and uncertainty, Appl. Therm. Eng. 56 (2013) 491e504.
134e151. [37] A.-J.N. Khalifa, Natural convective heat transfer coefficient e a review: II.
[6] ASHRAE, Handbook e Fundamentals, American Society of Heating, Refriger- Surfaces in two- and three-dimensional enclosures, Energy Convers. Manage.
ating, and Air-Conditioning Engineers, Inc., Atlanta, USA, 1985. 42 (2001) 505e517.
[7] TRANSSOLAR Energietechnik GmbH, Multizone building modeling with [38] F. Allard, Contribution to the Study of Heat Transfer in Thermally Driven
Type56 and TRNBuild, in: TRNSYS 17 a TRaNsient SYstem Simulation Program, Cavity with Large Rayleigh Number: Application to Residential Units (Ph.D.
Madison, USA, vol. 5, 2010, p. 56. thesis), Institut National des Sciences Applique es, Lyon, France, 1987 (in
[8] Q. Chen, C.A. Mayers, J. van der Kooi, Convective heat transfer in rooms with French).
mixed convection, in: International Seminar on Indoor Air Flow Patterns, [39] C. Martin, M. Watson, Measurement of Convective Heat Transfer Coefficients
University of Liege, Lie
ge, Belgium, 1989, pp. 69e82. in a Realistic Room Geometry (Technical Report), Energy Monitoring Com-
[9] H.B. Awbi, A. Hatton, Natural convection from heated room surfaces, Energy pany, Ltd, London, UK, 1988.
Build. 30 (1999) 233e244. [40] J. Gao, J. Zhang, J.N. Zhao, Numerical determination of convection coefficients
[10] A.J.N. Khalifa, R.H. Marshall, Natural and forced convection on interior for internal surfaces in buildings dominated by thermally stratified flows,
building surfaces-preliminary results, in: Proceedings of the Applied Energy J. Build. Phys. 31 (2008) 213e223.
Research Conference, 5e7 September 1989, University of Swansea, Swansea, [41] F. Causone, S.P. Corgnati, M. Filippi, B.W. Olesen, Experimental evaluation of
UK, 1989, pp. 249e257. heat transfer coefficients between radiant ceiling and room, Energy Build. 41
[11] A.J.N. Khalifa, R.H. Marshall, Validation of heat transfer coefficients on interior (2009) 622e628.
building surfaces using a real-sized indoor test cell, Int. J. Heat Mass Transfer [42] L. Peeters, I. Beausoleil-Morrison, A. Novoselac, Internal convective heat
33 (1990) 2219e2236. transfer modeling: critical review and discussion of experimentally derived
[12] A.J.N. Khalifa, Heat Transfer Processes in Buildings (PhD thesis), University of correlations, Energy Build. 43 (2011) 2227e2239.
Wales College of Cardiff, Cardiff, UK, 1989. [43] S. Leenknegt, R. Wagemakers, W. Bosschaerts, D. Saelens, Numerical sensi-
[13] T.C. Min, L.F. Schutrum, G.V. Parmelee, J.D. Vouris, Natural convection and tivity study of transient surface convection during night cooling, Energy Build.
radiation in a panel-heated room, ASHRAE Trans. 62 (1956) 337e358. 53 (2012) 85e95.
[14] Solar Energy Laboratory, Mathematical reference, in: TRNSYS 17 a TRaNsient [44] S.A. Al-Sanea, M.F. Zedan, M.B. Al-Harbi, Effect of supply Reynolds number
SYstem Simulation Program, vol. 4, University of Wisconsin-Madison, Madi- and room aspect ratio on flow and ceiling heat-transfer coefficient for mixing
son, USA, 2010, pp. 336e337. ventilation, Int. J. Therm. Sci. 54 (2012) 176e187.
[15] A. Hatton, H.B. Awbi, Convective heat transfer in rooms, in: Building Simu- [45] A. Baïri, Correlations for transient natural convection in parallelogrammic
lation '95, Fourth International Conference Proceedings, Madison, USA, 1995, enclosures with isothermal hot wall, Appl. Therm. Eng. 51 (2013) 833e838.
pp. 221e227. [46] C.R. Ruivo, P.M. Ferreira, D.C. Vaz, Prediction of thermal load temperature
[16] F. Alamdari, G.P. Hammond, Improved data correlations for buoyancy-driven difference values for the external envelope of rooms with setback and setup
convection in rooms, Build. Serv. Eng. Res. Technol. 4 (1983) 106e112. thermostats, Appl. Therm. Eng. 51 (2013) 980e987.
[17] G.F.C. Rogers, Y.R. Mayhew, Engineering Thermodynamics: Work and Heat [47] C. Yan, S. Wang, K. Shan, Y. Lu, A simplified analytical model to evaluate the
Transfer, Longmans, London, UK, 1967. impact of radiant heat on building cooling load, Appl. Therm. Eng. 77 (2015)
[18] ASHRAE, Handbook of Fundamentals, Chartered Institution of Building Ser- 30e41.
vices Engineers (CIBSE), Atlanta, USA, 1981. [48] J.A. Palyvos, A survey of wind convection coefficient correlations for building
[19] CIBSE Guide C3, Heat Transfer, Chartered Institution of Building Services En- envelope energy systems' modelling, Appl. Therm. Eng. 28 (2008) 801e808.
gineers (CIBSE), London, UK, 1976. [49] A. Novoselac, B.J. Burley, J. Srebric, Development of new and validation of
[20] R. Karadag, New approach relevant to total heat transfer coefficient including existing convection correlations for rooms with displacement ventilation
the effect of radiation and convection at the ceiling in a cooled ceiling room, systems, Energy Build. 38 (2006) 163e173.
Appl. Therm. Eng. 29 (2009) 1561e1565. [50] C. Massart, A. De Herde, Development of a Tool to Help in the Design of
[21] D. Schlapmann, Convection with Floor Heating-Development of a Test Houses at Very Low Energy Consumption Sustainable New Homes, Architec-
Method, Report No. BMFT FB-T, Universitat Stuttgart, Stuttgart, Germany, ture et Climat e UCL, Service Public Wallonie, Belgium, 2010 (in French).
1981, pp. 81e158. [51] A. Liebard, A. De Herde, Treaty of Bioclimatic Architecture and Urbanism,
[22] L.D. Li, W.A. Beckman, J.W. Mitchell, An Experimental Study of Natural Con- 
Editions du moniteur ed., 2005. Le Moniteur/Observ'ER, (in French).
vection in an Office Room, Large Time Results (Unpublished Report), Solar [52] CIBSE, Guide F: Energy Efficiency in Buildings, second ed., Chartered Institu-
Energy Laboratory, University of Wisconsin, Madison, USA, 1983. tion of Building Services Engineers (CIBSE), Norwich, UK, 2004.
[23] S. Fohanno, G. Polidori, Modelling of natural convective heat transfer at an [53] CIBSE Guide A, Chartered Institution of Building Services Engineers (CIBSE),
internal surface, Energy Build. 38 (2006) 548e553. seventh ed., Chartered Institution of Building Services Engineers, London, UK,
[24] G.N. Walton, Thermal Analysis Research Program Reference Manual, NBSSIR 2006.
83-2655, National Bureau of Standards, Washington, USA, 1983. [54] ASHRAE, ASHRAE Handbook e Fundamentals (S-I Edition), American Society
[25] W.W. Yousef, J.D. Tarasuk, W.J. McKeen, Free convection heat transfer from of Heating, Refrigerating and Air-Conditioning Engineers, Inc, Atlanta, USA,
upward-facing isothermal horizontal surfaces, J. Heat Transfer 104 (1982) 2009.
493e500. [55] W.J. King, The Basic Laws and Data of Heat Transmission, American Society of
[26] D.E. Fisher, An Experimental Investigation of Mixed Convection Heat Transfer Mechanical Engineers, New York, USA, 1932.
in a Rectangular Enclosure (Ph.D. thesis), University of Illinois, Urbana- [56] W.H. McAdams, Heat Transmission, McGraw-Hill, London, UK, 1954 (Printed
Champaign, USA, 1995. in New-York, U.S.A.).
[27] D.E. Fisher, C.O. Pedersen, Convective heat transfer in building energy and [57] O.A. Saunders, M. Fishenden, H.D. Mansion, Some measurements of convec-
thermal load calculations, ASHRAE Trans., Boston, USA 103 (1997) 137e148. tion by an optical method, Engineering (1935) 483e485.
[28] K. Goldstein, A. Novoselac, Convective heat transfer in rooms with ceiling slot [58] M. Al-Arabi, M.K. El-Riedy, Natural convection heat transfer from isothermal
diffusers (RP-1416), HVAC R Res. 16 (2010) 629e655. horizontal plates of different shapes, Int. J. Heat Mass Transfer 19 (1976)
[29] I. Beausoleil-Morrison, The Adaptive Coupling of Heat and Air Flow Modeling 1399e1404.
Within Dynamic Whole-building Simulations (PhD. thesis), University of [59] R.C.L. Bosworth, Heat Transfer Phenomena; the Flow of Heat in Physical Systems,
Strathclyde, Glasgow, UK, 2000. Associated General Publications, J. Wiley, Sydney, New York, USA, 1952.
[30] I. Beausoleil-Morrison, An algorithm for calculating convection coefficients for [60] R. Ishiguro, Heat transfer and flow instability of natural convection over
internal building surfaces for the case of mixed flow in rooms, Energy Build. upward-facing horizontal surfaces, in: Proceedings of the Sixth International
33 (2001) 351e361. Heat Transfer Conference, NC-8, Toronto, Canada, 1978, pp. 229e334.
[31] H.B. Awbi, A. Hatton, Mixed convection from heated room surfaces, Energy [61] E. Griffiths, A.H. Davis, The Transmission of Heat by Radiation and Convection,
Build. 32 (2000) 153e166. Dept. of Scientific and Industrial Research, Food Investigation Board, Engi-
[32] S.W. Churchill, R. Usagi, A general expression for the correlation of rates of neering Committee, H. M. Stationery Off., London, UK, 1922.
transfer and other phenomena, AIChE J. 18 (1972) 1121e1128. [62] T. Aihara, Y. Yamada, S. Endo €, Free convection along the downward-facing
[33] H. Akbari, D. Samano, A. Mertol, F. Bauman, R. Kammerud, The effect of var- surface of a heated horizontal plate, Int. J. Heat Mass Transfer 15 (1972)
iations in convection coefficients on thermal energy storage in buildings part I 2535e2549.
e interior partition walls, Energy Build. 9 (1986) 195e211. [63] F. Restrepo, L.R. Glicksman, The effect of edge conditions on natural convec-
[34] I. Beausoleil-Morrison, Flow responsive modelling of internal surface con- tion from a horizontal plate, Int. J. Heat Mass Transfer 17 (1974) 135e142.
vection, in: Seventh International IBPSA Conference. Rio de Janeiro, Brazil, [64] M.W. Fishenden, O.A. Saunders, An Introduction to Heat Transfer, Clarendon
2001, pp. 923e930. Press, Oxford, UK, 1950.

You might also like