You are on page 1of 8

Energy and Buildings 63 (2013) 59–66

Contents lists available at SciVerse ScienceDirect

Energy and Buildings


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

A computational investigation of a room heated by subcutaneous


convection—A case study of a replica Roman bath
Taylor Oetelaar a,∗ , Clifton Johnston a,b , David Wood a , Lisa Hughes c , John Humphrey c
a
Department of Mechanical and Manufacturing Engineering, University of Calgary, 2500 University Dr NW, Calgary, AB, Canada T2N 1N4
b
Department of Mechanical Engineering, Dalhousie University, 5269 Morris Street, Halifax, NS, Canada B3H 4R2
c
Department of Greek and Roman Studies, University of Calgary, 2500 University Dr NW, Calgary, AB, Canada T2N 1N4

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

Article history: Floor and wall subcutaneous convective heating is a common and efficient supplementary system of
Received 1 May 2012 heating, ventilation, and air conditioning (HVAC) but the concept dates back nearly 2500 years to when
Received in revised form 25 March 2013 ancient Greeks and Romans used it to warm their bathing facilities. This paper explores the thermal envi-
Accepted 26 March 2013
ronment of a replica Roman bath resulting from purely subcutaneous convective heating by modelling
the bath using computational fluid dynamics (CFD). Previous studies examining the interior of baths have
Keywords:
used either heat fluxes or lumped-mass thermodynamics, but neither approach possesses the detail of
Floor heating
CFD. The average temperature in the 3 m × 4 m room modelled is 35 ◦ C; however, the hottest air is trapped
Wall heating
Thermal comfort
in the high vault leaving the region inhabited by the patrons significantly cooler than the average. The
Computational fluid dynamics results also show that stratification is prominent and that the open doorway connecting to the next room
Roman baths heavily influences the room temperature. The results also suggest a relative insensitivity to changes in
the convective heat transfer coefficient and addition of humidity to the model. Furthermore, this study
not only provides further knowledge about an alternative HVAC system but enhances our understanding
of ancient Roman baths. In addition, it offers an insight to a unique thermal environment on the basis of
multi-phase and species modelling.
© 2013 Elsevier B.V. All rights reserved.

1. Introduction system. The case study is the hot bath room or caldarium of a replica
of an ancient Roman bath built for the PBS television series NOVA
Most modern HVAC systems rely on forced air to bring in by a team led by Yegül and Couch [3] located outside Sardis, Turkey.
fresh air and heat or cool homes and buildings [1]. There has The CFD package that we used was FLUENT 6.3/ANSYS FLUENT 13.0.
been a move recently towards the use of in-floor ‘radiant slab’ By comparison with most Roman baths, the building is relatively
heating to aid the standard HVAC forced air systems [1]. While small (Fig. 1b), measuring 7.6 m long by 6.4 m wide by 2.8 m high
the use of in-floor heating is often seen as a modern advance- to the base of the vaulting, and has three rooms. The room of inter-
ment, the premise actually dates back nearly twenty-five hundred est, known as the caldarium, or hot bath room, is in the southwest
years. The ancient Roman Empire made extensive use of subcu- corner of the building. Its floor area is 2.84 m by 3.02 m and the
taneous convective floor heating in their public and private baths air volume is just over 42 m3 . The east wall has an open doorway
[2]. The premise behind the Roman heating system, known as the which connects to the next room called the tepidarium or ‘room
hypocaust (Fig. 1a), was simple but extremely effective. A fire in with warm baths’ and has a cloth door. The south wall has an alcove
a furnace (praefurnia) creates hot exhaust gases which circulate for a hot pool. The north, west, and east walls have the tubuli up to
between the foundation floor and a suspended floor (suspensura) just below the springing of the vault; these, in combination with
and up box tubes (tubuli) inlaid in the walls. The exhaust gases give the floor, supply the heat for the room. Finally, there are three small
off heat to the room before escaping to the atmosphere through windows, totalling 2 m2 , above the tubuli in the west wall.
flues. Our work uses CFD to analyze the thermal environment, set-
ting aside air quality issues, of a room heated using only a hypocaust 2. Literature

The application of CFD to analyze buildings is rare because of the


∗ Corresponding author.
long simulation time often required. The majority of analyses are
E-mail addresses: taylor.a.oetelaar@gmail.com (T. Oetelaar),
conducted using building simulation. Most CFD models are of sin-
clifton.johnston@dal.ca (C. Johnston), dhwood@ucalgary.ca (D. Wood), gle rooms, such as offices [4–8], or generic rooms [9–11] to analyze
lahughes@ucalgary.ca (L. Hughes), humphrey@ucalgary.ca (J. Humphrey). thermal comfort, which is a significant aspect of HVAC design. There

0378-7788/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.enbuild.2013.03.049
60 T. Oetelaar et al. / Energy and Buildings 63 (2013) 59–66

performed a very detailed sun study of the Terme del Foro at Ostia
and determined that it would have been possible to comfortably
bathe nude in the open rooms without glass in the windows. Ring
[26], using thermodynamics, concluded that it was not feasible to
rely on the heat from the sun and the hypocaust to reach an accept-
able temperature without the use of glazing. Jorio [27] provides
detailed information about the hypocausts of the Pompeian baths
(Stabian, Forum, and Central) and then analyzed the heat losses
from four rooms (the men’s and women’s caldaria and tepidaria) in
the Stabian Baths. Rook [28] utilized thermodynamic equations to
estimate the fuel consumed by the Welwyn Roman bath as 13 kg/h
or 114 tonnes/yr. Yegül and Couch [3] also took measurements of
various parameters when they ran their replica–many of which we
used in our simulations–and performed a heat loss analysis. All of
these studies, though, deal in heat fluxes or average temperatures,
which do not provide as detailed an analysis as CFD. CFD has the
ability to show how the temperature is distributed and the loca-
tions of any drafts and hot spots that could dramatically affect our
understanding and interpretation of caldaria and how people used
them.

3. Methodology

3.1. Grid

We chose to start with an unstructured mesh combining tetra-


hedral and hexahedral elements since the air volume is quite
complex in shape. The initial mesh was 1,976,183 cells with no
cells above 80% skewness, meaning all cells were close to their
ideal shape. This mesh was used to do preliminary set up, includ-
ing contrasting turbulence models, density models and temporal
dependency. A structured mesh was developed prior to includ-
ing the effects of the heated pool. To ensure continuity, the results
from a structured mesh without the pool were compared with the
Fig. 1. (a) Diagram of the hypocaust heating system. (b) Digital model of the NOVA results from the unstructured mesh. After establishing congruency
baths from the southwest looking northeast. between the results from the two mesh types, we performed a grid
refinement test with different cell sizes (100 mm, 50 mm, 37.5 mm,
and 25 mm with a corresponding change in the total number of
are some important exceptions. Some studies investigated ther-
cells). We found that grid independence was attained at a cell size
mal patterns and/or air quality issues in large open spaces [12–18].
of 37.5 mm, with minor fluctuations near the window. Therefore,
Balocco’s analysis [19] of the Salone dei Duecento di Palazzo Vec-
the final structured mesh used 37.5 mm cells with 10 mm cells near
chio (Hall of the Two Hundred of the Old Palace) in Florence
the window and had 990,143 cells without the pool and 1,164,501
modelled the interior environment to help with tapestry preser-
cells with the pool.
vation. Ayata and Yildiz [20] proposed a new way of incorporating
natural ventilation for buildings in Turkey through optimal place-
ment of windows in the façade. Stavrakakis [21] demonstrated the 3.2. Sub-models
impact of cross-ventilation on thermal comfort by simulating a
small building with two doors. While these studies focused on mod- In this simulation there are five important sub-models: time-
ern HVAC systems and environments, they all are predicated on the dependency, density, turbulence, multi-phase, and species. The
use of, at least in part, forced air to control temperature. Our case first run was done as steady-state but subsequent runs were time-
study predates, technologically-speaking, the advent of electricity dependent because this stabilized the multi-phase calculations.
and, as such, the thermal environment will necessarily be different. The effect of time dependency appeared to be dramatic because
In terms of bathing complexes, the use of numerical methods is the direction of the air velocity near the doorway to the tepidar-
very uncommon and Basaran has led the studies that have used this ium reversed when the model was changed. Further investigation
approach. In the first, Basaran and Ilken [22] did not use CFD but revealed this as an anomaly for the reasons described below.
rather numerical heat transfer to investigate the heating system Three density models were compared: the Boussinesq approx-
of the Small Baths at Phaselis in Turkey. Their simulations sug- imation, the incompressible ideal gas law, and ideal gas law. The
gested a number of key conclusions, including: (1) the bath only results of the Boussinesq approximation proved very dependent on
had limited use in the winter; (2) the input heat was less than the the reference temperature. If the reference temperature was below
heat lost; (3) the fuel required was massive. In the second, Basaran the bulk temperatures of the room the flow patterns in the room
et al. [23] then used CFD to analyze the heating system of a bath varied drastically from those of the two other models. Therefore,
house in Metropolis and found that the hypocaust design was inef- we rejected the Boussinesq approximation was rejected because
ficient and the fuel requirements were substantial. The third and of this sensitivity to reference temperature. There were no differ-
final [24], is, in essence, a summary of the first two and again reit- ences in any parameter between the results of the two gas laws. We
erates the inefficiency of the heating system and high fuel costs. chose the incompressible ideal gas law because it did not require
There are, however, a number of analytical studies. Thatcher [25] additional pressure calculations, which can slow the simulation.
T. Oetelaar et al. / Energy and Buildings 63 (2013) 59–66 61

Table 1
The boundary conditions for the wall sections.

Property Glazing External wall Heated floor Heated wall Connecting wall

Type Mixed Mixed Convective Convective Convective


Material Glass Custom Custom Custom Custom
Thickness (m) 0.005 0.40 0.20 0.06 0.40
Convective heat transfer coefficient (W/m2 K) 4 4 7 7 4
Free stream temperature (K) 288 288 363 363 300
External emissivity 0.49 0.94 – – –
External radiation temperature (K) 0 323 – – –

Four turbulence models were tested: RNG k–ε, SST k–T, and (the lower 2 m of the room height). The results of Yegül and Couch
those two with viscous heating. (In these designations, k represents [3] agreed with the predicted warmer temperature. As a result, we
the turbulent kinetic energy, ε is the turbulent dissipation rate, ω adopted the SST k–T model, however, either model was suitable.
is the specific dissipation rate, RNG indicates a renormalized group Only the implicit volume of fluid (VOF) and the “Transport &
theory modification, and SST stands for shear stress transport. For Reaction” models were applicable for multi-phase and species sim-
more information, refer to Wilcox [29].) We compared results with ulation, respectively. All other multi-phase and species models did
and without viscous heating for both turbulence models and the not meet our requirements. However, the difficult part is modelling
effect of viscous heating proved negligible. The difference between the evaporation of the water from the pool to the air. We used a
the RNG k-, and SST k–T was relatively small with the SST k–T pro- user-defined function (UDF) to accomplish this. The UDF first iden-
ducing slightly warmer results (within 3 ◦ C) in the occupied region tifies the interface between the two phases where the evaporation

Fig. 2. (a) The temperature distribution in the x-mid-plane for Case #1 (units: ◦ C). (b) The temperature distribution in the z-mid-plane for Case #1 (units: ◦ C). (c) The velocity
vectors in the x-mid-plane for Case #1 (units: m/s). (d) The velocity vectors in the z-mid-plane for Case #1 (units: m/s).
62 T. Oetelaar et al. / Energy and Buildings 63 (2013) 59–66

Table 2
Case breakdown.

Case Time-dependency Turbulence Multi-phase Species Heated wall CHTC (W/m2 K) Radiation model

#1 – Base Unsteady SST k–ω N/A N/A 2 N/A


#2 Unsteady SST k–ω VOF S&T 2 N/A
#3 Unsteady SST k–ω VOF S&T 7 N/A
#4 Unsteady SST k–ω VOF S&T 7 S2S

is taking place through an area function and then calculates the with no thermal significance a zero heat flux (insulating) condition.
mass transfer rate on this interface. We also tested the effect of The conditions for the remaining walls are summarized in Table 1.
including FLUENT’s surface-to-surface (S2S) radiation model. For the wall connected to the tepidarium the most suitable is the
convective boundary condition. This is because the heat transfer
is due to the natural convection of the air in the tepidarium and
3.3. Boundary conditions the fact there is no separate radiation heat source. Once this type
has been selected, the next question concerns the convective heat
In this simulation there are a number of key zones: the exterior transfer coefficient (CHTC). Since the wall is relatively large it can
walls, the window glass, the heated floor, the heated walls, the wall be assumed to behave similarly to a vertical plate, on which the
connecting to the tepidarium, the doorway to the tepidarium, the behaviour of the CHTC is well-known [30]. For a vertical plate under
water inlet to the pool, the water outlet, and the walls that do not normal room conditions, a good first estimate is 4 W/m2 K [31]. The
have thermal significance (i.e., the door jamb). We gave the zones emissivity for the glass came from a FLUENT tutorial [31] on HVAC

Fig. 3. (a) The temperature distribution in the x-mid-plane for the Case #2 (units: ◦ C). (b) The temperature distribution in the z-mid-plane for the Case #2 (units: ◦ C). (c) The
velocity vectors in the x-mid-plane for the Case #2 (units: m/s). (d) The velocity vectors in the z-mid-plane for the Case #2 (units: m/s).
T. Oetelaar et al. / Energy and Buildings 63 (2013) 59–66 63

testing and we retrieved the emissivity for the external wall from
the entry for rough concrete on the Engineering Toolbox website
[32].
The heated walls pose an interesting problem. The most appro-
priate boundary condition is a convective one since that is the
principle behind the design–hot air rising from below the sub-floor
and out the chimneys. The temperature of the free stream is not dif-
ficult to estimate as Yegül and Couch give various readings of the air
inside the system [3]. The question then is what the CHTC should
be. For preliminary computer runs a value of 2 W/m2 K was used
as the flues were confined spaces which might indicate lower val-
ues than normal. However, for the penultimate simulation a more
accurate assessment was necessary. The best way of determining
this value was through experimentation. Based on these results
[33], the CHTC was set to 7 W/m2 K. The free stream temperatures
came from the data in Ref. [3] and the external radiation tempera-
ture came from Ref. [31]. We used the maximum recorded flue gas
temperature for the heated surfaces.
The water inlets and outlets are uncomplicated. The water
comes in at 313 K and an arbitrary low velocity of 0.1 m/s and
the backflow temperature of the outlet is 312 K. This is a common
temperature for many hot tub tests [34].

4. Results and discussion


Fig. 4. The water vapour distribution in the z-mid-plane for the Case #2 (units:
There are three important cases to demonstrate the progres- kg/kg).
sion of the simulation and are designated in Table 2. Displaying
meaningful results from a 3-D model is somewhat difficult as one
can only take flat slices or use iso-surfaces. For economy, only two comes into the room further. This flow reversal completely changes
slices were chosen to display the temperature distributions and the environment within the room. The entire room is colder, par-
velocity vectors: the mid-planes of both the x- and z-directions. ticularly the occupied region, and the velocities are higher in the
These slices were chosen because they capture the major features room. This pattern though is more likely than the one from Case #1
of the NOVA bath—the hypocausted walls, the hypocausted floor, because of the more accurate modelling. The heating system has
the windows, the pool, and the doorway to the adjoining room—and the same effect: it creates an updraft and a layer of warm air next
they represent the three Cartesian directions. to the walls.
The temperature distribution and velocity vector profile of the There are two new smaller drafts in Case #2. One from the win-
Case #1 are shown in Fig. 2a–d. Overall, the temperature ranges dows, which is due to an adjustment in the solar load calculator.
from 27 ◦ C to 38 ◦ C and the velocity magnitudes are low with Inadvertently, Case #1 did not include a feature that spread the
most of the room being largely stagnant. The first thing that is solar heating to adjacent cells. Without this feature enabled, the
immediately apparent upon closer inspection of the temperature solar heat would create unrealistic hot spots. We corrected this
distributions is how, without a fan to stir the air, buoyancy drives oversight in Case #2. The more significant draft is the one originat-
the movement as the stratification is pronounced. From a comfort ing from the pool. In Case #1, the stand-in for the pool did not affect
perspective, however, this stratification has a drawback. It means the flow regime greatly, but the addition of the water warmed the
that the region occupied by the patrons (i.e., the volume of air below alcove by 10 ◦ C and impeded the flow of cool air from the tepidar-
the height of the door or 1.88 m) is only between 30 ◦ C and 35 ◦ C, ium. However, the temperature effects of the pool remains localized
which is only slightly warmer than the tepidarium, which is a warm to the alcove, further showing the influence of the doorway.
room. Therefore, while the average temperature of the room might The humidity profile in Fig. 4 illustrates the effect of the pool
be 30 ◦ C, much of the heat from the hypocaust is lost to the vault. well. There is a definite stream of water vapour rising from the
The other significant factor seen from both the temperature and water/air interface with the vapour culminating at the apex of the
velocity vector profiles is the impact of the doorway. The cooler alcove vault before dispersing into the room. The humidity itself
air from the tepidarium does not fully enter the room and thus it does not have a noticeable effect on the room temperature. This
has a limited effect on the overall temperature. The flow pattern is not surprising since the room is dominated by the tepidarium
is unique and almost counter-intuitive because it is coming in the door boundary condition which means most of the room is at the
top and leaving the bottom. Moving away from the doorway, there tepidarium humidity.
is a thin film of warm air above the floor and next to the heated The only change with Case #3 was that the convective heat
walls showing the effect of these surfaces. The pool stand-in does transfer coefficient (CHTC) of the heated walls was raised from the
not appear to have much impact on the overall temperature distri- early estimate of 2 W/m2 K to the experimentally derived value
bution but there is a draft coming off the surface which indicates of 7 W/m2 K [33]. The results (Fig. 5a–d) illustrate the resulting
that it is perturbing the flow. changes. A comparison of Figs. 3b and 5b illustrates an increase
With the addition of humidity and the pool the flow pattern in temperature along the heated wall from the 30–35 ◦ C to the
changes dramatically as seen in Fig. 3a–d. It all, however, stems 35–40 ◦ C, indicating that the increased CHTC is generating more
from the reversal of the doorway flow. In Case #1, the air from the heat. There is a similar trend with the z-mid-plane temperature
tepidarium came in from the top and exited out the bottom thereby distribution (Fig 5b). The cooling effect of the air coming in from
limiting the effect of its cooler nature. In Case #2, though, the air the tepidarium is significantly reduced. These changes result in
enters at the bottom and exits at the top, which means the cooler air the room being generally warmer with the experimentally derived
64 T. Oetelaar et al. / Energy and Buildings 63 (2013) 59–66

Fig. 5. (a) The temperature distribution in the x-mid-plane for Case #3 (units: ◦ C). (b) The temperature distribution in the z-mid-plane for Case #3 (units: ◦ C). (c) The velocity
vectors in the x-mid-plane for Case #3 (units: m/s). (d) The velocity vectors in the z-mid-plane for Case #3 (units: m/s).

CHTC. However, the temperature has not increased significantly Part of this decreased response to CHTC is possibly due to the
in proportion to the increase in the CHTC. The value of the CHTC insulating nature of the wall material. Because of its lower ther-
increased 350% but the volume-averaged temperature in the room mal conductivity, a ceramic wall such as this requires more heat to
only increased 2 ◦ C (32.8–34.8 ◦ C), which is less than seven percent. make a set temperature increase than a metallic or modern com-
This suggests that while the CHTC of the heated walls is a drives the posite wall. This could be by design as a safety feature, but is more
simulation the temperature distribution in the room is relatively likely an inherent benefit of the materials available to Romans. With
insensitive to changes in CHTC. In fact, the heat from the heated 90 ◦ C exhaust gases only 6 cm from the surface of the wall, some-
walls only increases by between 139% and 189% with an average of one leaning against it could be seriously burned without this type
168% as seen in Table 3. of wall.

Table 3
Comparison of heat fluxes calculated by FLUENT.

Zone Case #2 input heat (W) Case #3 input heat (W) % Increase #2 to #3 Case #4 input heat (W) % Increase #3 to #4

Tubuli next to doorway 51.1 96.6 189.0 151.9 157.2


North Tubuli 460.7 787.4 170.9 1032.5 131.1
NE Tubuli 149.9 253.6 169.2 313.9 123.8
SE Tubuli 170.7 283.3 166.0 392.4 138.5
W Tubuli 535.6 934.0 174.4 1172.4 125.5
Top of Tubuli 70.2 97.6 139.0 195.2 200.0

Average 168.1 146.0


T. Oetelaar et al. / Energy and Buildings 63 (2013) 59–66 65

The velocity vector profiles are very similar. The velocities in


Case #3 are slightly lower than the other cases. However, this is
only a marginal change. The main currents are almost identical in
the two cases.
The water vapour (Fig. 6) reaches a higher concentration in the
corner of the alcove above the pool in the humidity distribution in
Case #3. It does not reach as far into the main room. This change is
due to the minor alteration in the air current above the pool.
The addition of the radiation model in Case #4 (Fig. 7a–d) does
not affect the results significantly, especially in the occupied region.
The average temperature by volume only differs by approximately
0.1 ◦ C (34.8 ◦ C for Case #3 versus 34.7 ◦ C for Case #4). The flow pat-
terns are similar with the only important change at the top of the
vault where the temperature breaches the next temperature con-
tour. Also, Table 3 shows that the change in heat fluxes is less than
between Cases #2 and #3. It is also interesting to note that the
largest increase in heat flux came from the top of the tubuli; that is,
above where the people are. The major zones only rose an average
of 30% and, while substantial, this, again, did not affect the over-
all temperature of the room. Based on these data, we believe that
the change resulting from addition of radiation model is negligible.
However, given the increase in heat flux, there might be a percep-
tual change in the thermal environment (i.e., it might feel warmer)
but this is beyond the scope of the paper.
Fig. 6. The water vapour distribution in the z-mid-plane for Case #3 (units: kg/kg).
Case #3 appears to compare well to the data from Yegül and
Couch [3]. The average temperature of the caldarium given is 35 ◦ C

Fig. 7. (a) The temperature distribution in the x-mid-plane for Case #4 (units: ◦ C). (b) The temperature distribution in the z-mid-plane for Case #4 (units: ◦ C). (c) The velocity
vectors in the x-mid-plane for Case #4 (units: m/s). (d) The velocity vectors in the z-mid-plane for Case #4 (units: m/s).
66 T. Oetelaar et al. / Energy and Buildings 63 (2013) 59–66

and the volume-averaged temperature from the computational for a typical office building in Singapore, Energy and Buildings 37 (2005)
model is 34.8 ◦ C. However, one potential limitation of our simu- 563–572.
[7] Z. Lin, T.T. Chow, C.S. Csang, K.F. Fong, L.S. Chan, Effects of headroom on the per-
lation is that we used the maximum temperature Yegul and Couch formance of the displacement ventilation system effects of headroom on the
listed for the exhaust gases. We feel that this is necessitated by our performance of the displacement ventilation system, Indoor and Built Environ-
consideration of the doorway to the tepidarium. We have modelled ment 15 (4) (2006) 333–346.
[8] J.A. Myhren, S. Holmberg, Flow patterns and thermal comfort in a room with
the door as an outlet with a constant temperature of 27 ◦ C based on panel, floor and wall heating, Energy and Buildings 40 (2008) 524–536.
the tepidarium temperature reported by Yegül and Couch. In actu- [9] T. Kim, S. Kato, S. Murakami, Indoor cooling/heating load analysis based on
ality, the temperature across the length of the two rooms will vary coupled simulation of convection, radiation and HVAC control, Building and
Environment 36 (2001) 901–908.
since there is no wall separating them. The exchange between these [10] C. Teodosi, R. Hohota, G. Rusaouën, M. Woloszyn, Numerical prediction of
two rooms should be gradual. In fact, Yegül states that he designed indoor air humidity and its effect on indoor environment, Building and Envi-
the baths so that the caldarium would heat, at least partially, the ronment 38 (2003) 655–664.
[11] C.M. Mak, J.L. Niu, C.T. Lee, K.F. Chan, A numerical simulation of wing walls using
tepidarium. Nonetheless these three case studies still illustrate the
computational fluid dynamics, Energy and Buildings 39 (2007) 995–1002.
effectiveness of the heating system as well as the impact, or lack [12] K.A. Papakonstantinou, C.T. Kiranoudis, N.C. Markatos, Numerical simulation
there of, of the addition of humidity. of CO2 dispersion in an auditorium, Energy and Buildings 34 (2002) 245–250.
[13] M. Kavgic, D. Mumovic, Z. Stevanovic, A. Young, Analysis of thermal comfort and
indoor air quality in a mechanically ventilated theatre, Energy and Buildings
5. Conclusions 40 (2008) 1334–1343.
[14] K.A. Papakonstantinou, C.T. Kiranoudis, N.C. Markatos, Computational analysis
In this study, we have modelled a room heated by a subcuta- of thermal comfort: the case of the archaeological museum of Athens, Applied
Mathematical Modelling 24 (2000) 477–494.
neous convective heating system rather than the modern mixing [15] A.I. Stamou, I. Katsiris, A. Schaelin, Evaluation of thermal comfort in Galatsi
of air through a simulation of a room inside a modern replica of Arena of the Olympics “Athens 2004” using a CFD model, Applied Thermal
an ancient Roman bath built for the television series NOVA. With- Engineering 28 (2008) 1206–1215.
[16] K.D. Song, Evaluating daylighting and heating designs of a top-glazed atrium
out the constant influx of new air though ventilation, stratification space through physical scale model measurements and CFD analyses, Indoor
becomes prominent and any opening in the envelope of the room and Built Environment 16 (2) (2007) 121–129.
becomes extremely important to the air circulation patterns. In [17] J. Lau, J.L. Niu, Measurement, CFD simulation of the temperature stratification in
an atrium using a floor level air supply method, Indoor and Built Environment
this case, the doorway to the tepidarium cools the occupied region 12 (2003) 265–280.
considerably which negates the heating system. The difference [18] Y. Lin, R. Zmeureanu, Computer model of the airflow and thermal phenomena
between a wall with the same thermal properties as water and a inside a large dome, Energy and Buildings 40 (2008) 1287–1296.
[19] C. Balocco, Daily natural heat convection in a historical hall, Journal of Cultural
simulated pool is remarkable. The pool had a much greater impact
Heritage 8 (2007) 370–376.
on the flow by creating a draft of warmer air. Interestingly, how- [20] T. Ayata, O. Yildiz, Investigating the potential use of natural ventilation in new
ever, the CHTC of the heated walls and floor has little impact on building designs in Turkey, Energy and Buildings 38 (2006) 959–963.
[21] G.M. Stavrakakis, M.K. Koukou, M.G. Vrachopoulos, N.C. Markatos, Natural
the overall temperature. The volume-averaged temperature rose
cross-ventilation in buildings: building-scale experiments, numerical sim-
by 2 ◦ C even though the CHTC increased by three hundred and fifty ulation and thermal comfort evaluation, Energy and Buildings 40 (2008)
percent from 2 W/m2 K to 7 W/m2 K. Finally, this study provides 1666–1681.
a methodological starting point for future CFD studies on Roman [22] T. Basaran, Z. Ilken, Thermal analysis of the heating system of the small bath in
ancient phaselis, Energy and Buildings 27 (1998) 1–11.
baths. As the first application of CFD to the interior of a Roman [23] T. Basaran, A. Erek, G. Uluans, A. Ersoy, Energy analysis of the roman bath
bath, we have shown that CFD can highlight previously untested in metropolis, in: The Second International Exergy, Energy and Environment
aspects. Heat flux analyses and lumped-mass thermodynamics, the Symposium, July 3–7, Kos Island, Greece, 2005.
[24] T. Basaran, The heating system of the Roman Baths, ASHRAE Transactions 113
previous approaches, for example, do not take into account the (2007) 199–205.
stratification due to buoyancy which has a tremendous impact [25] E. Thatcher, Open rooms at the Terme del Foro at Ostia, Memoirs of the Amer-
on the thermal environment of the room. This methodology may ican Academy in Rome 24 (1956) 169–264.
[26] J.W. Ring, Windows, baths, and solar energy in the Roman empire, American
stimulate different perspectives in terms of architectural or social Journal of Archaeology 100 (1996) 717–724.
studies. [27] A. Jorio, Sistema di Riscaldamento nelle Antiche Terme Pompeiane, Bullettino
Della Commissione Archeologica Comunale Di Roma 86 (1978–1979) 167–189.
[28] T. Rook, The development and operation of Roman hypocausted baths, Journal
References of Archaeological Science 5 (1978) 269–282.
[29] C.D. Wilcox, Turbulence Modeling for CFD, second ed., DCW Industries, La
[1] K. Moe, Thermally Active Surfaces in Architecture, Princeton Architectural Cañada, 1998.
Press, New York, 2010. [30] J.P. Holman, Heat Transfer, ninth ed., McGraw Hill, Boston, 2002.
[2] F.K. Yegul, Baths and Bathing in Classical Antiquity, The MIT Press, Cambridge, [31] Fluent Inc., Tutorial: Using Solar Load Model for Indoor Ventilation, Fluent Inc.,
1995. Canonsburg, PA, Tutorial, 2007.
[3] F.K. Yegül, T. Couch, Building a Roman bath for the cameras, Journal of Roman [32] Engineering Toolbox, Emissivity Coefficients of Some Common Materi-
Archaeology 16 (2003) 153–177. als, http://www.engineeringtoolbox.com/emissivity-coefficients-d 447.html,
[4] H. Manz, T. Frank, Analysis of thermal comfort near cold vertical surfaces by Retrieved April 19, 2012.
means of computational fluid dynamics, Indoor and Built Environment 13 [33] T.A. Oetelaar, C.R. Johnston, Determination of the convective heat transfer coef-
(2004) 233–242. ficient of hot air rising through terracotta flues, Transactions of the Canadian
[5] J. Abantoa, D. Barreroa, M. Reggioa, B. Ozella, Air flow modelling in a computer Society for Mechanical Engineering 36 (4) (2012) 413–428.
room, Building and Environment 39 (2004) 1393–1402. [34] M.A.S. Harvey, M.M. McRorie, D.W. Smith, Suggested limits to the use of the
[6] W.N. Hien, W. Liping, A.N. Chandra, A.R. Pandey, W. Xiaolin, Effects of dou- hot tub and sauna by pregnant women, Canadian Medical Association Journal
ble glazed facade on energy consumption, thermal comfort and condensation 125 (1981) 50–53.

You might also like