Você está na página 1de 8

Energy and Buildings 63 (2013) 5966

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 convectionA case study of a replica Roman bath
Taylor Oetelaar a, , Clifton Johnston a,b , David Wood a , Lisa Hughes c , John Humphrey c
a b c

Department of Mechanical and Manufacturing Engineering, University of Calgary, 2500 University Dr NW, Calgary, AB, Canada T2N 1N4 Department of Mechanical Engineering, Dalhousie University, 5269 Morris Street, Halifax, NS, Canada B3H 4R2 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
Floor and wall subcutaneous convective heating is a common and efcient supplementary system of heating, ventilation, and air conditioning (HVAC) but the concept dates back nearly 2500 years to when ancient Greeks and Romans used it to warm their bathing facilities. This paper explores the thermal environment of a replica Roman bath resulting from purely subcutaneous convective heating by modelling the bath using computational uid dynamics (CFD). Previous studies examining the interior of baths have used either heat uxes or lumped-mass thermodynamics, but neither approach possesses the detail of CFD. The average temperature in the 3 m 4 m room modelled is 35 C; however, the hottest air is trapped in the high vault leaving the region inhabited by the patrons signicantly cooler than the average. The results also show that stratication is prominent and that the open doorway connecting to the next room heavily inuences the room temperature. The results also suggest a relative insensitivity to changes in the convective heat transfer coefcient 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.

Article history: Received 1 May 2012 Received in revised form 25 March 2013 Accepted 26 March 2013 Keywords: Floor heating Wall heating Thermal comfort Computational uid dynamics Roman baths

1. Introduction Most modern HVAC systems rely on forced air to bring in fresh air and heat or cool homes and buildings [1]. There has been a move recently towards the use of in-oor radiant slab heating to aid the standard HVAC forced air systems [1]. While the use of in-oor heating is often seen as a modern advancement, the premise actually dates back nearly twenty-ve hundred years. The ancient Roman Empire made extensive use of subcutaneous convective oor heating in their public and private baths [2]. The premise behind the Roman heating system, known as the hypocaust (Fig. 1a), was simple but extremely effective. A re in a furnace (praefurnia) creates hot exhaust gases which circulate between the foundation oor and a suspended oor (suspensura) and up box tubes (tubuli) inlaid in the walls. The exhaust gases give off heat to the room before escaping to the atmosphere through ues. Our work uses CFD to analyze the thermal environment, setting aside air quality issues, of a room heated using only a hypocaust

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 by a team led by Yegl and Couch [3] located outside Sardis, Turkey. The CFD package that we used was FLUENT 6.3/ANSYS FLUENT 13.0. By comparison with most Roman baths, the building is relatively small (Fig. 1b), measuring 7.6 m long by 6.4 m wide by 2.8 m high to the base of the vaulting, and has three rooms. The room of interest, known as the caldarium, or hot bath room, is in the southwest corner of the building. Its oor area is 2.84 m by 3.02 m and the air volume is just over 42 m3 . The east wall has an open doorway which connects to the next room called the tepidarium or room with warm baths and has a cloth door. The south wall has an alcove for a hot pool. The north, west, and east walls have the tubuli up to just below the springing of the vault; these, in combination with the oor, supply the heat for the room. Finally, there are three small windows, totalling 2 m2 , above the tubuli in the west wall. 2. Literature The application of CFD to analyze buildings is rare because of the long simulation time often required. The majority of analyses are conducted using building simulation. Most CFD models are of single rooms, such as ofces [48], or generic rooms [911] to analyze thermal comfort, which is a signicant aspect of HVAC design. There

Corresponding author. E-mail addresses: taylor.a.oetelaar@gmail.com (T. Oetelaar), clifton.johnston@dal.ca (C. Johnston), dhwood@ucalgary.ca (D. Wood), lahughes@ucalgary.ca (L. Hughes), humphrey@ucalgary.ca (J. Humphrey). 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) 5966

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 acceptable 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 mens and womens 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. Yegl and Couch [3] also took measurements of various parameters when they ran their replicamany of which we used in our simulationsand performed a heat loss analysis. All of these studies, though, deal in heat uxes 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 locations 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 tetrahedral 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, including contrasting turbulence models, density models and temporal dependency. A structured mesh was developed prior to including the effects of the heated pool. To ensure continuity, the results from a structured mesh without the pool were compared with the results from the unstructured mesh. After establishing congruency between the results from the two mesh types, we performed a grid renement test with different cell sizes (100 mm, 50 mm, 37.5 mm, and 25 mm with a corresponding change in the total number of cells). We found that grid independence was attained at a cell size of 37.5 mm, with minor uctuations near the window. Therefore, the nal structured mesh used 37.5 mm cells with 10 mm cells near the window and had 990,143 cells without the pool and 1,164,501 cells with the pool. 3.2. Sub-models In this simulation there are ve important sub-models: timedependency, density, turbulence, multi-phase, and species. The rst run was done as steady-state but subsequent runs were timedependent because this stabilized the multi-phase calculations. The effect of time dependency appeared to be dramatic because the direction of the air velocity near the doorway to the tepidarium reversed when the model was changed. Further investigation revealed this as an anomaly for the reasons described below. Three density models were compared: the Boussinesq approximation, the incompressible ideal gas law, and ideal gas law. The results of the Boussinesq approximation proved very dependent on the reference temperature. If the reference temperature was below the bulk temperatures of the room the ow patterns in the room varied drastically from those of the two other models. Therefore, we rejected the Boussinesq approximation was rejected because of this sensitivity to reference temperature. There were no differences in any parameter between the results of the two gas laws. We chose the incompressible ideal gas law because it did not require additional pressure calculations, which can slow the simulation.

Fig. 1. (a) Diagram of the hypocaust heating system. (b) Digital model of the NOVA baths from the southwest looking northeast.

are some important exceptions. Some studies investigated thermal patterns and/or air quality issues in large open spaces [1218]. Baloccos analysis [19] of the Salone dei Duecento di Palazzo Vecchio (Hall of the Two Hundred of the Old Palace) in Florence modelled the interior environment to help with tapestry preservation. Ayata and Yildiz [20] proposed a new way of incorporating natural ventilation for buildings in Turkey through optimal placement of windows in the fac ade. Stavrakakis [21] demonstrated the impact of cross-ventilation on thermal comfort by simulating a small building with two doors. While these studies focused on modern HVAC systems and environments, they all are predicated on the use of, at least in part, forced air to control temperature. Our case study predates, technologically-speaking, the advent of electricity and, as such, the thermal environment will necessarily be different. In terms of bathing complexes, the use of numerical methods is very uncommon and Basaran has led the studies that have used this approach. In the rst, Basaran and Ilken [22] did not use CFD but rather numerical heat transfer to investigate the heating system of the Small Baths at Phaselis in Turkey. Their simulations suggested a number of key conclusions, including: (1) the bath only had limited use in the winter; (2) the input heat was less than the heat lost; (3) the fuel required was massive. In the second, Basaran et al. [23] then used CFD to analyze the heating system of a bath house in Metropolis and found that the hypocaust design was inefcient and the fuel requirements were substantial. The third and nal [24], is, in essence, a summary of the rst two and again reiterates the inefciency of the heating system and high fuel costs. There are, however, a number of analytical studies. Thatcher [25]

T. Oetelaar et al. / Energy and Buildings 63 (2013) 5966 Table 1 The boundary conditions for the wall sections. Property Type Material Thickness (m) Convective heat transfer coefcient (W/m2 K) Free stream temperature (K) External emissivity External radiation temperature (K) Glazing Mixed Glass 0.005 4 288 0.49 0 External wall Mixed Custom 0.40 4 288 0.94 323 Heated oor Convective Custom 0.20 7 363 Heated wall Convective Custom 0.06 7 363

61

Connecting wall Convective Custom 0.40 4 300

Four turbulence models were tested: RNG k, SST kT, and those two with viscous heating. (In these designations, k represents the turbulent kinetic energy, is the turbulent dissipation rate, is the specic dissipation rate, RNG indicates a renormalized group theory modication, and SST stands for shear stress transport. For more information, refer to Wilcox [29].) We compared results with and without viscous heating for both turbulence models and the effect of viscous heating proved negligible. The difference between the RNG k-, and SST kT was relatively small with the SST kT producing slightly warmer results (within 3 C) in the occupied region

(the lower 2 m of the room height). The results of Yegl and Couch [3] agreed with the predicted warmer temperature. As a result, we adopted the SST kT model, however, either model was suitable. Only the implicit volume of uid (VOF) and the Transport & Reaction models were applicable for multi-phase and species simulation, respectively. All other multi-phase and species models did not meet our requirements. However, the difcult part is modelling the evaporation of the water from the pool to the air. We used a user-dened function (UDF) to accomplish this. The UDF rst identies 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 Table 2 Case breakdown. Case #1 Base #2 #3 #4 Time-dependency Unsteady Unsteady Unsteady Unsteady

T. Oetelaar et al. / Energy and Buildings 63 (2013) 5966

Turbulence SST k SST k SST k SST k

Multi-phase N/A VOF VOF VOF

Species N/A S&T S&T S&T

Heated wall CHTC (W/m2 K) 2 2 7 7

Radiation model N/A N/A N/A S2S

is taking place through an area function and then calculates the mass transfer rate on this interface. We also tested the effect of including FLUENTs surface-to-surface (S2S) radiation model.

3.3. Boundary conditions In this simulation there are a number of key zones: the exterior walls, the window glass, the heated oor, the heated walls, the wall connecting to the tepidarium, the doorway to the tepidarium, the water inlet to the pool, the water outlet, and the walls that do not have thermal signicance (i.e., the door jamb). We gave the zones

with no thermal signicance a zero heat ux (insulating) condition. The conditions for the remaining walls are summarized in Table 1. 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 the fact there is no separate radiation heat source. Once this type has been selected, the next question concerns the convective heat transfer coefcient (CHTC). Since the wall is relatively large it can be assumed to behave similarly to a vertical plate, on which the behaviour of the CHTC is well-known [30]. For a vertical plate under normal room conditions, a good rst estimate is 4 W/m2 K [31]. The 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) 5966

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 appropriate boundary condition is a convective one since that is the principle behind the designhot air rising from below the sub-oor and out the chimneys. The temperature of the free stream is not difcult to estimate as Yegl 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 ues were conned spaces which might indicate lower values 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 temperature came from Ref. [31]. We used the maximum recorded ue 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 backow temperature of the outlet is 312 K. This is a common temperature for many hot tub tests [34].

4. Results and discussion There are three important cases to demonstrate the progression of the simulation and are designated in Table 2. Displaying meaningful results from a 3-D model is somewhat difcult as one can only take at slices or use iso-surfaces. For economy, only two slices were chosen to display the temperature distributions and velocity vectors: the mid-planes of both the x- and z-directions. These slices were chosen because they capture the major features of the NOVA baththe hypocausted walls, the hypocausted oor, the windows, the pool, and the doorway to the adjoining roomand they represent the three Cartesian directions. The temperature distribution and velocity vector prole of the Case #1 are shown in Fig. 2ad. Overall, the temperature ranges from 27 C to 38 C and the velocity magnitudes are low with most of the room being largely stagnant. The rst thing that is immediately apparent upon closer inspection of the temperature distributions is how, without a fan to stir the air, buoyancy drives the movement as the stratication is pronounced. From a comfort perspective, however, this stratication has a drawback. It means that the region occupied by the patrons (i.e., the volume of air below the height of the door or 1.88 m) is only between 30 C and 35 C, which is only slightly warmer than the tepidarium, which is a warm room. Therefore, while the average temperature of the room might be 30 C, much of the heat from the hypocaust is lost to the vault. The other signicant factor seen from both the temperature and velocity vector proles is the impact of the doorway. The cooler air from the tepidarium does not fully enter the room and thus it has a limited effect on the overall temperature. The ow pattern is unique and almost counter-intuitive because it is coming in the top and leaving the bottom. Moving away from the doorway, there is a thin lm of warm air above the oor and next to the heated walls showing the effect of these surfaces. The pool stand-in does not appear to have much impact on the overall temperature distribution but there is a draft coming off the surface which indicates that it is perturbing the ow. With the addition of humidity and the pool the ow pattern changes dramatically as seen in Fig. 3ad. It all, however, stems from the reversal of the doorway ow. In Case #1, the air from the tepidarium came in from the top and exited out the bottom thereby limiting the effect of its cooler nature. In Case #2, though, the air enters at the bottom and exits at the top, which means the cooler air
Fig. 4. The water vapour distribution in the z-mid-plane for the Case #2 (units: kg/kg).

comes into the room further. This ow reversal completely changes the environment within the room. The entire room is colder, particularly the occupied region, and the velocities are higher in the room. This pattern though is more likely than the one from Case #1 because of the more accurate modelling. The heating system has the same effect: it creates an updraft and a layer of warm air next to the walls. There are two new smaller drafts in Case #2. One from the windows, which is due to an adjustment in the solar load calculator. Inadvertently, Case #1 did not include a feature that spread the solar heating to adjacent cells. Without this feature enabled, the solar heat would create unrealistic hot spots. We corrected this oversight in Case #2. The more signicant draft is the one originating from the pool. In Case #1, the stand-in for the pool did not affect the ow regime greatly, but the addition of the water warmed the alcove by 10 C and impeded the ow of cool air from the tepidarium. However, the temperature effects of the pool remains localized to the alcove, further showing the inuence of the doorway. The humidity prole in Fig. 4 illustrates the effect of the pool well. There is a denite stream of water vapour rising from the water/air interface with the vapour culminating at the apex of the alcove vault before dispersing into the room. The humidity itself does not have a noticeable effect on the room temperature. This is not surprising since the room is dominated by the tepidarium door boundary condition which means most of the room is at the tepidarium humidity. The only change with Case #3 was that the convective heat transfer coefcient (CHTC) of the heated walls was raised from the early estimate of 2 W/m2 K to the experimentally derived value of 7 W/m2 K [33]. The results (Fig. 5ad) illustrate the resulting changes. A comparison of Figs. 3b and 5b illustrates an increase in temperature along the heated wall from the 3035 C to the 3540 C, indicating that the increased CHTC is generating more heat. There is a similar trend with the z-mid-plane temperature distribution (Fig 5b). The cooling effect of the air coming in from the tepidarium is signicantly reduced. These changes result in the room being generally warmer with the experimentally derived

64

T. Oetelaar et al. / Energy and Buildings 63 (2013) 5966

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 signicantly in proportion to the increase in the CHTC. The value of the CHTC increased 350% but the volume-averaged temperature in the room only increased 2 C (32.834.8 C), which is less than seven percent. This suggests that while the CHTC of the heated walls is a drives the simulation the temperature distribution in the room is relatively insensitive to changes in CHTC. In fact, the heat from the heated walls only increases by between 139% and 189% with an average of 168% as seen in Table 3.

Part of this decreased response to CHTC is possibly due to the insulating nature of the wall material. Because of its lower thermal conductivity, a ceramic wall such as this requires more heat to make a set temperature increase than a metallic or modern composite wall. This could be by design as a safety feature, but is more likely an inherent benet of the materials available to Romans. With 90 C exhaust gases only 6 cm from the surface of the wall, someone leaning against it could be seriously burned without this type of wall.

Table 3 Comparison of heat uxes calculated by FLUENT. Zone Tubuli next to doorway North Tubuli NE Tubuli SE Tubuli W Tubuli Top of Tubuli Average Case #2 input heat (W) 51.1 460.7 149.9 170.7 535.6 70.2 Case #3 input heat (W) 96.6 787.4 253.6 283.3 934.0 97.6 % Increase #2 to #3 189.0 170.9 169.2 166.0 174.4 139.0 168.1 Case #4 input heat (W) 151.9 1032.5 313.9 392.4 1172.4 195.2 % Increase #3 to #4 157.2 131.1 123.8 138.5 125.5 200.0 146.0

T. Oetelaar et al. / Energy and Buildings 63 (2013) 5966

65

Fig. 6. The water vapour distribution in the z-mid-plane for Case #3 (units: kg/kg).

The velocity vector proles 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. 7ad) does not affect the results signicantly, 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 ow patterns are similar with the only important change at the top of the vault where the temperature breaches the next temperature contour. Also, Table 3 shows that the change in heat uxes is less than between Cases #2 and #3. It is also interesting to note that the largest increase in heat ux 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 overall 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 ux, there might be a perceptual change in the thermal environment (i.e., it might feel warmer) but this is beyond the scope of the paper. Case #3 appears to compare well to the data from Yegl 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) 5966 for a typical ofce building in Singapore, Energy and Buildings 37 (2005) 563572. Z. Lin, T.T. Chow, C.S. Csang, K.F. Fong, L.S. Chan, Effects of headroom on the performance of the displacement ventilation system effects of headroom on the performance of the displacement ventilation system, Indoor and Built Environment 15 (4) (2006) 333346. J.A. Myhren, S. Holmberg, Flow patterns and thermal comfort in a room with panel, oor and wall heating, Energy and Buildings 40 (2008) 524536. T. Kim, S. Kato, S. Murakami, Indoor cooling/heating load analysis based on coupled simulation of convection, radiation and HVAC control, Building and Environment 36 (2001) 901908. C. Teodosi, R. Hohota, G. Rusaoun, M. Woloszyn, Numerical prediction of indoor air humidity and its effect on indoor environment, Building and Environment 38 (2003) 655664. C.M. Mak, J.L. Niu, C.T. Lee, K.F. Chan, A numerical simulation of wing walls using computational uid dynamics, Energy and Buildings 39 (2007) 9951002. K.A. Papakonstantinou, C.T. Kiranoudis, N.C. Markatos, Numerical simulation of CO2 dispersion in an auditorium, Energy and Buildings 34 (2002) 245250. M. Kavgic, D. Mumovic, Z. Stevanovic, A. Young, Analysis of thermal comfort and indoor air quality in a mechanically ventilated theatre, Energy and Buildings 40 (2008) 13341343. K.A. Papakonstantinou, C.T. Kiranoudis, N.C. Markatos, Computational analysis of thermal comfort: the case of the archaeological museum of Athens, Applied Mathematical Modelling 24 (2000) 477494. A.I. Stamou, I. Katsiris, A. Schaelin, Evaluation of thermal comfort in Galatsi Arena of the Olympics Athens 2004 using a CFD model, Applied Thermal Engineering 28 (2008) 12061215. K.D. Song, Evaluating daylighting and heating designs of a top-glazed atrium space through physical scale model measurements and CFD analyses, Indoor and Built Environment 16 (2) (2007) 121129. J. Lau, J.L. Niu, Measurement, CFD simulation of the temperature stratication in an atrium using a oor level air supply method, Indoor and Built Environment 12 (2003) 265280. Y. Lin, R. Zmeureanu, Computer model of the airow and thermal phenomena inside a large dome, Energy and Buildings 40 (2008) 12871296. C. Balocco, Daily natural heat convection in a historical hall, Journal of Cultural Heritage 8 (2007) 370376. T. Ayata, O. Yildiz, Investigating the potential use of natural ventilation in new building designs in Turkey, Energy and Buildings 38 (2006) 959963. G.M. Stavrakakis, M.K. Koukou, M.G. Vrachopoulos, N.C. Markatos, Natural cross-ventilation in buildings: building-scale experiments, numerical simulation and thermal comfort evaluation, Energy and Buildings 40 (2008) 16661681. T. Basaran, Z. Ilken, Thermal analysis of the heating system of the small bath in ancient phaselis, Energy and Buildings 27 (1998) 111. T. Basaran, A. Erek, G. Uluans, A. Ersoy, Energy analysis of the roman bath in metropolis, in: The Second International Exergy, Energy and Environment Symposium, July 37, Kos Island, Greece, 2005. T. Basaran, The heating system of the Roman Baths, ASHRAE Transactions 113 (2007) 199205. E. Thatcher, Open rooms at the Terme del Foro at Ostia, Memoirs of the American Academy in Rome 24 (1956) 169264. J.W. Ring, Windows, baths, and solar energy in the Roman empire, American Journal of Archaeology 100 (1996) 717724. A. Jorio, Sistema di Riscaldamento nelle Antiche Terme Pompeiane, Bullettino Della Commissione Archeologica Comunale Di Roma 86 (19781979) 167189. T. Rook, The development and operation of Roman hypocausted baths, Journal of Archaeological Science 5 (1978) 269282. C.D. Wilcox, Turbulence Modeling for CFD, second ed., DCW Industries, La Canada, 1998. J.P. Holman, Heat Transfer, ninth ed., McGraw Hill, Boston, 2002. Fluent Inc., Tutorial: Using Solar Load Model for Indoor Ventilation, Fluent Inc., Canonsburg, PA, Tutorial, 2007. Engineering Toolbox, Emissivity Coefcients of Some Common Materials, http://www.engineeringtoolbox.com/emissivity-coefcients-d 447.html, Retrieved April 19, 2012. T.A. Oetelaar, C.R. Johnston, Determination of the convective heat transfer coefcient of hot air rising through terracotta ues, Transactions of the Canadian Society for Mechanical Engineering 36 (4) (2012) 413428. M.A.S. Harvey, M.M. McRorie, D.W. Smith, Suggested limits to the use of the hot tub and sauna by pregnant women, Canadian Medical Association Journal 125 (1981) 5053.

and the volume-averaged temperature from the computational model is 34.8 C. However, one potential limitation of our simulation is that we used the maximum temperature Yegul and Couch listed for the exhaust gases. We feel that this is necessitated by our consideration of the doorway to the tepidarium. We have modelled the door as an outlet with a constant temperature of 27 C based on the tepidarium temperature reported by Yegl and Couch. In actuality, the temperature across the length of the two rooms will vary since there is no wall separating them. The exchange between these two rooms should be gradual. In fact, Yegl states that he designed the baths so that the caldarium would heat, at least partially, the tepidarium. Nonetheless these three case studies still illustrate the effectiveness of the heating system as well as the impact, or lack there of, of the addition of humidity. 5. Conclusions

[7]

[8] [9]

[10]

[11] [12] [13]

[14]

In this study, we have modelled a room heated by a subcutaneous convective heating system rather than the modern mixing of air through a simulation of a room inside a modern replica of an ancient Roman bath built for the television series NOVA. Without the constant inux of new air though ventilation, stratication becomes prominent and any opening in the envelope of the room becomes extremely important to the air circulation patterns. In this case, the doorway to the tepidarium cools the occupied region considerably which negates the heating system. The difference between a wall with the same thermal properties as water and a simulated pool is remarkable. The pool had a much greater impact on the ow by creating a draft of warmer air. Interestingly, however, the CHTC of the heated walls and oor has little impact on the overall temperature. The volume-averaged temperature rose by 2 C even though the CHTC increased by three hundred and fty percent from 2 W/m2 K to 7 W/m2 K. Finally, this study provides a methodological starting point for future CFD studies on Roman baths. As the rst application of CFD to the interior of a Roman bath, we have shown that CFD can highlight previously untested aspects. Heat ux analyses and lumped-mass thermodynamics, the previous approaches, for example, do not take into account the stratication due to buoyancy which has a tremendous impact on the thermal environment of the room. This methodology may stimulate different perspectives in terms of architectural or social studies. References
[1] K. Moe, Thermally Active Surfaces in Architecture, Princeton Architectural Press, New York, 2010. [2] F.K. Yegul, Baths and Bathing in Classical Antiquity, The MIT Press, Cambridge, 1995. [3] F.K. Yegl, T. Couch, Building a Roman bath for the cameras, Journal of Roman Archaeology 16 (2003) 153177. [4] H. Manz, T. Frank, Analysis of thermal comfort near cold vertical surfaces by means of computational uid dynamics, Indoor and Built Environment 13 (2004) 233242. [5] J. Abantoa, D. Barreroa, M. Reggioa, B. Ozella, Air ow modelling in a computer room, Building and Environment 39 (2004) 13931402. [6] W.N. Hien, W. Liping, A.N. Chandra, A.R. Pandey, W. Xiaolin, Effects of double glazed facade on energy consumption, thermal comfort and condensation

[15]

[16]

[17]

[18] [19] [20] [21]

[22] [23]

[24] [25] [26] [27] [28] [29] [30] [31] [32]

[33]

[34]

Você também pode gostar