You are on page 1of 15

Theor. Appl. Climatol. 87, 223237 (2007) DOI 10.

1007/s00704-005-0194-4

Meteorological Institute, University of Freiburg, Germany

Thermal comfort in an eastwest oriented street canyon in Freiburg (Germany) under hot summer conditions
F. Ali-Toudert and H. Mayer
With 10 Figures Received March 21, 2005; revised July 29, 2005; accepted November 4, 2005 Published online April 27, 2006 # Springer-Verlag 2006

Summary Field-measurements were conducted in an urban street canyon with an eastwest orientation, and a height-to-width ratio H=W 1 during cloudless summer weather in 2003 in Freiburg, Germany. This experimental work adds to the knowledge available on the microclimate of an urban canyon and its impact on human comfort. Air temperature Ta, air humidity VP, wind speed v and direction dd were measured continuously. All short-wave and long-wave radiation uxes from the 3D surroundings were also measured. The degree of comfort was dened in terms of physiologically equivalent temperature (PET). Furthermore, the data gathered within the canyon were compared to data collected by a permanent urban climate station with the aim of furthering the understanding of microclimatic changes due to street geometry. Changes in the meteorological variables Ta, v and dd in the canyon in comparison to an unobstructed roof level location were found to be in good agreement with previous studies, i.e., a small increase of Ta in the canyon adjacent to irradiated surfaces, and a good correlation of v and dd between canyon and roof levels. The daily dynamics of canyon facet irradiances and their impacts on the heat gained by a pedestrian were strongly dependent on street geometry and orientation. Thermal stress was mostly attributable to solar exposure. Under cloudless summer weather, a standing body was found to absorb, on average, 74% of heat in the form of long-wave irradiance and 26% as short-wave irradiance. Shading the pedestrian as well as the surrounding surfaces is, hence, the rst strategy in mitigating heat stress in summer under hot conditions.

1. Introduction With respect to urban climate, the urban street canyon is commonly considered as the basic structural unit of the urban canopy layer (e.g. Oke, 1988; Arneld, 1990). Basic knowledge of street climate has been provided by several studies which include thermal and energetic characteristics, air ow and air pollution (e.g. Nunez and Oke, 1977; Hussain and Lee, 1980; Oke, 1981, 1988; de Paul and Shieh, 1986; Nakamura and Oke, 1988; Arneld, 1990; Todhunter, 1990; Yoshida et al., 1990=91; Eliasson, 1993; Arneld and Mills, 1994; Sini et al., 1996; Kim and Baik, 1999; Santamouris et al., 1999; Uehara et al., 2000; Coronel and Alvarez, 2001). In contrast, only a limited number of studies have been undertaken to quantify the effects of the urban thermal environment on human comfort. Whereas most of these studies have focused on urban land use differences (e.g. Mayer and Hppe, 1987; Mayer, 1993; Svensson et al., o 2003), or behavioural aspects (e.g. Nagara et al., 1996; Nikolopoulou et al., 2001; Spagnolo and de Dear, 2003a, 2003b; Stathopoulos et al., 2004; Nikolopoulou and Lykoudis, 2005), studies directly addressing the role of street geometry, which are explicitly design-oriented, still remain

224

F. Ali-Toudert and H. Mayer

scarce (e.g. Swaid et al., 1993; Pearlmutter et al., 1999; Grundstrm et al., 2003; Ali-Toudert, o 2005; Ali-Toudert and Mayer, 2006; Ali-Toudert et al., 2005). On both issues of street microclimate and outdoor comfort, there has been a greater tendency to use numerical modelling methods rather to conduct experimental studies. The popularity of numerical modelling, over the last decades (Arneld, 2003), is largely attributable to the costly and time consuming exercise of directly recording all the relevant meteorological variables using accurate measurement methods. In particular, continuous observations of radiation uxes surrounding a human body in open spaces are lacking. For convenience, this is commonly replaced by a globe thermometer as an integral instrument (e.g. Nikolopoulou et al., 2001; Nikolopoulou and Lykoudis, 2005), even though it is only accurate indoors (ASHRAE, 2001). Yet, in order to validate results obtained from the modelling of urban microclimate, there is need to collect extensive data (Arneld, 2003). In this study, therefore, emphasis is placed on the experimental method, which is based on a comprehensive measure of all variables inuencing human outdoor comfort with analysis of their dependence on urban geometry. The results include an analysis of the street microclimate and, more crucially, the radiant environment induced by canyon geometry and their incidence on the thermal sensation of a standing person. 2. Site and measurements In-situ measurements were conducted in an urban street canyon (Erbprinzenstrae) in the downtown area of Freiburg (48 000 N, 7 500 E and 280 m a.s.l.), a medium-sized city in the southern upper Rhine plain in southwest Germany, on the western slopes of the Black Forest mountainous region. The canyon axis is oriented in an eastwest direction (Fig. 1). The street is anked by long buildings, which preserve the canyon alignment for at least 150 m, despite the presence of a number of gaps in the building fronts. At the measuring site, the canyon is symmetric with an aspect ratio H=W 1 and a sky view factor SVF 0.26 (Fig. 2).

Fig. 1. Plan view of the eastwest oriented urban street canyon street in Freiburg, Germany, showing the location of the street station at the northern sidewalk and the measuring points MP1 to MP4

Fig. 2. Fish-eye photography of the urban street canyon in Freiburg, Germany, at the location of the street station at northern sidewalk (top of Fig. is N wall)

The buildings are almost of equal height, typically of two or three stories with pitched tile roofs. The street is made of asphalt and is 12 m wide. The walls of the buildings are constructed of bricks and are painted with light colours. Windows constitute about 30% of the walls. A small park with tall trees is located in the vicinity of the canyon on the west side. The east end of the canyon opens onto a small planted area while the western end connects to a main northsouth orientated road. Some sparse vegetation along the street is also present. The experimental work was conducted on 14 and 15 July 2003: two hot, sunny days. Although the period of data collection was short, the prevailing conditions on these two days were considered representative of a typical hot summer in Freiburg.

Thermal comfort in an eastwest oriented street canyon under hot summer conditions

225

Table 1. Meteorological instrumentation used at the station located in an EW urban street canyon in Freiburg, Germany Variable Short-wave radiation Long-wave radiation Net all-wave radiation Air temperature Air humidity Wind speed Wind direction Instrument Pyranometer (CM11, CM21, Kipp & Zonen) difference between net all-wave and short-wave radiation Pyradiometer (Schenk) Pt-100 (HMP Vaisala) Humicap (HMP Vaisala) cup anemometer (Vector Instruments) wind vane (Vector Instruments) Height (a.g.l.) 1.4 m 1.4 m 1.4 m 1.4 m, 3.1 m 1.4 m, 3.1 m 1.4 m, 3.1 m 3.1 m Number 6 (", #, N, E, S, W) 6 (", #, N, E, S, W) 6 (", #, N, E, S, W) 2 2 2 1

A vertical mast tted with temperature, wind and radiation sensors was installed at a distance of 1 m from the northern wall (i.e. station in Fig. 1) and referred to as street station. This location corresponds to the pedestrian sidewalk and hence, thermal comfort is required. This is also the most critical location with respect to comfort for this orientation (EW), as reported by a previous study (Ali-Toudert and Mayer, 2006). Air temperature, air humidity, wind speed and wind direction were continuously recorded at regular time intervals at two heights: 1.4 m and 3.1 m a.g.l. (above ground level). The short-wave and long-wave radiation ux densities were measured from the three-dimensional surroundings, i.e. upwards and downwards together with the four lateral directions (N, E, S and W). All meteorological variables were recorded in the form of 10-minute-averages (scan interval: 10 s) over a 30-hour-period. The instrumentation used at the street station is listed in Table 1 and the site is illustrated in Fig. 2. In order to obtain a spatially differentiated picture of the street microclimate, supplementary readings were collected on 14 July and included manually taken measurements of air and surface temperatures on both sides of the street (MP1 to MP4, Fig. 1). In addition, the data obtained in the street were compared to those provided by a permanent urban climate station in order to clarify the microclimatic changes within the canyon. This is referred to as background station and is run by the Meteorological Institute, University of Freiburg (MIF, 2005). The background station is located on the roof of a high-rise building at a height of 51 m a.g.l. It is situated in the northern part of Freiburg, approximately 1500 m from the investigated street. Air temperature and humidity sensors were placed 2 m above roof level (a.r.l.) according to the psychrometer method, as well as

the radiation sensor (global radiation). The wind sensors (cup anemometer and wind vane) were placed at 10 m a.r.l. The data of the background station and the street station were compared in order to clarify the microclimatic changes inside the canyon due to the obstructing effects of the buildings. 3. Thermal comfort assessment Thermophysiologically signicant indices have been developed in human-biometeorology to assess the human thermal comfort outdoors, e.g., predicted mean vote PMV (Jendritzky et al., 1990), outdoor standard temperature OUT-SET (Pickup and de Dear, 1999), physiologically signicant temperature PET (Hppe, 1993, 1999) or pero ceived temperature PT (Staiger et al., 1997). Despite some detailed differences in the parameterizations employed, all these indices rely on the human energy balance and are applicable outdoors. In this study, PET has been used. By denition, PET is the air temperature at which, in a typical indoor setting (Tmrt Ta, VP 12 hPa, v 0.1 ms1 ), the heat balance of the human body, assuming light activity and a heat transfer resistance of the clothing of 0.9 clo, is maintained with core and skin temperature equal to those under actual conditions (Hppe, 1993). This ino dex is well documented and has been widely used in urban climate investigations (e.g. Mayer and Hppe, 1987; Mayer, 1993; Svensson et al., o 2003; Ali-Toudert and Mayer, 2006; Ali-Toudert et al., 2005). The mean radiant temperature Tmrt, which is rather difcult to determine outdoors with accuracy (Ali-Toudert, 2005) is required for the calculation of PET. Hppe (1992) described a o method to calculate Tmrt on the basis of measured short-wave and long-wave radiation ux densi-

226

F. Ali-Toudert and H. Mayer

ties from the three-dimensional surroundings of humans. The formula for Tmrt (in  C) is:   Srad 0:25 Tmrt 273:2 1 "p B with Srad
6 X i1

Wi ak Ki al Li

where (Ki) and (Li) (i 1; . . . ; 6) represent, respectively, the short-wave and the long-wave radiation ux densities in the six directions; the total radiation ux density is denoted Srad. The angle factors Wi are the percentage of these uxes received by the human body in each direction. For a standing person (assimilated to a cylinder-like shape) Wi equals 0.22 for lateral directions and 0.06 upwards and downwards. The short-wave and long-wave absorption coefcients are denoted ak and al and assume the values 0.7 and 0.97, respectively. The emissivity of the human body "p takes the value 0.97 and the Stefan-Boltzman constant: B 5.67 108 Wm2 K2 . 4. Results 4.1. Microclimate within the urban street canyon 4.1.1 Air and surface temperatures Figure 3 shows the daily course of the air temperature Ta as recorded by the stations in the canyon and above-roof, together with the supplementary readings measured manually at the

four additional points along the sidewalks on 14 July. The air temperature Ta within the urban street canyon varied between 18  C and 35  C indicating the relatively high thermal level during the measuring period (Fig. 3), part of the recordbreaking heat-wave which affected Europe during the summer of 2003 (Fink et al., 2004). In the canyon, there was little difference in Ta measured at various points before 1300 LST and after 1800 LST owing to the well mixed air inside the canyon. During the afternoon of 14 July, from 1400 to 1800 LST, Ta measured at the sunlit part of the street, i.e. at the street station, MP3 and MP4, was a few degrees higher than those recorded on the opposite side of the street (MP1 and MP2) which were mostly in shade. Compared to Ta above roof level, almost no difference is found during the period between 0800 and 1300 LST, during which time the street is yet to warm up. In the afternoon, Ta was higher at the street station as well as at MP3 and MP4, than above roof level. At night, the street is cooler than the free air above roof (at 53 m a.g.l.) with a maximum difference Ta of 3 K. This feature is apparently anomalous and could be attributed to the microclimatic differences within the city of Freiburg reported by Nbler (1979) and to the inversion u forming above the mean roof level of Freiburg (approximately 20 m a.g.l.). The urban canyon surfaces facing south were irradiated during most of the day and experienced high ground surface temperature Ts and wall surface temperature Tw values (Fig. 4a, calculated from L with the emissivity " 0.98), leading to

Fig. 3. Daily variation of the air temperature Ta at two heights at the street station (hourly mean values) and at different measuring points (MP) within the EW oriented street canyon (H=W 1) as well as at the roof level of a high-rise building in Freiburg, Germany, on two typical summer days (14=15 July 2003)

Thermal comfort in an eastwest oriented street canyon under hot summer conditions

227

Fig. 4. Hourly mean values of (a) air temperature Ta, ground surface temperature Ts and wall surface temperature Tw at the street station on the northern sidewalk of the urban canyon (H=W 1) in Freiburg, Germany, on two typical summer days (14=15 July 2003) and (b) mean Ts and Tw values at the points MP1 and MP2 (southern sidewalk) and MP3 and MP4 (northern sidewalk) on a typical summer day (14 July 2003)

the increased transfer of heat to air as sensible heat ux. Figure 4a also shows that peak surface temperatures Ts and Tw occurred around 1500 LST which resulted in the air temperature Ta reaching its maximum one hour later (1600 LST). Comparing both sides of the street on Fig. 3, Ta at the south side (MP1 and MP2) showed lower values than at the north side (station, MP3 and MP4) due to the limited heat released by the adjacent surfaces, as these had noticeably lower temperatures (Fig. 4b). The maximum Ta difference between the two sidewalks was reached at 1700 LST and was approximately 3 K for the same height (1.4 m). At the street station, Ta at 1.4 m a.g.l. was higher than at 3.1 m a.g.l. as a consequence of the increased proximity to the sunlit ground and wall surfaces. The peak Ta difference reached 1.2 K in the afternoon, whereas it was negligible in the evening hours between all measuring points within the canyon, and did not exceed 0.5 K.

The results of Ta are in good agreement with previous studies conducted for streets with almost the same characteristics: EW orientation and H=W 1. Nakamura and Oke (1988) found that Ta shows small differences between roof air and canyon air, except close to sunlit urban facets where the heat transferred from the heated walls led to warmer adjacent air. Yoshida et al. (1990=91) conrmed the insignicant warming of canyon air in comparison to free ambient air and the relative homogeneity of Ta across and along a street canyon. They also reported on large differences in the surface temperatures between sunlit and shaded surfaces. Surfaces in shade are noticeably cooler than irradiated surfaces, and the surface temperatures can even be lower than Ta. Santamouris et al. (1999) reported almost similar ndings for a deeper street canyon oriented close to NS, but with moderate vertical thermal stratication. Air temperature differences of up to

228

F. Ali-Toudert and H. Mayer

Fig. 5. Ten-minutes mean values of (a) wind direction dd and (b) horizontal wind speed v within the urban street canyon (3.1 m a.g.l.) and above roof level (61 m a.g.l.) on two typical summer days (14=15 July 2003) in Freiburg, Germany

2 K were found between irradiated and shaded sidewalks in various urban canyons under hot summer conditions (e.g. Nakamura and Oke, 1988). 4.1.2 Wind direction and wind speed The relationship between wind ow above-roof and within the canyon is shown in Figs. 5 and 6. The following analysis should be read bearing in mind the following uncertainties: The background station is about 1500 m from the street station and the data describing the above-roof wind conditions were recorded at 61 m a.g.l. This is about four times the canyon height and as a result the wind speed directly at roof level (13 m a.g.l.) is much lower.

The air ow in the canyon is known to be a secondary circulation feature driven by the above-roof imposed ow (e.g. Nakamura and Oke, 1988; Santamouris et al., 1999). Basically, the correlation between the canyon wind and the wind above roof level is found to be more marked for high wind speeds, whereas the coupling between the upper and secondary ow is lost at lower velocities, leading to much more scattering (Figs. 5 and 6b). During the measuring period, the wind was either parallel or oblique. Almost no perpendicular incidence was recorded. Three distinct and temporarily consecutive episodes could be observed, with remarkably different combinations of wind directions and speeds, which have in turn inuenced the wind ow characteristics within the

Thermal comfort in an eastwest oriented street canyon under hot summer conditions

229

Fig. 6. (a) Wind speed v and (b) wind direction dd from the background station (61 m a.g.l.) plotted against corresponding values within the urban canyon (H=W 1) at 3.1 m a.g.l. on two typical summer days (14=15 July 2003) in Freiburg, Germany

canyon. The wind direction in the canyon depended on the angle of incidence in relation to the canyons axis of the upper wind (Fig. 5a). When the wind above-roof is nearly parallel to the canyon axis (30 ), the wind in the canyon ows in the same direction due to channelling (Nakamura and Oke, 1988; Santamouris et al., 1999). In this case, it corresponded to a thermally induced local circulation system known as Hllentler, which blows from the east (Black o a Forest mountainous region) with relatively high speed (Ernst, 1995). This occurred during the night between 2000 to 0630 LST. On the rst day, from 1200 LST to 2000 LST, the wind was blowing at an angle of incidence with moderate

velocity: from the NW quadrant and faster than 5 ms1 . This led to a wind inside the canyon owing from the SW direction. This ow scheme has been described as a spiral vortex induced along the canyon (e.g. Wedding et al., 1977; Nakamura and Oke, 1988; Santamouris et al., 1999). The simple relationship suggested by Nakamura and Oke (1988) for an urban canyon with H=W 1 applied for relatively high winds, that is ddcanyon 540 dd roof for dd roof 180 360, with dd 360 being the north, seems to apply to the present case study as a rst approximation. From 1000 to 1300 LST on 14 July and after 0630 LST on 15 July, weak winds with no dominant direction prevailed above roof level. This led to a large scattering in the canyon. In this case, the wind ow in the canyon was not only a mechanically driven circulation but thermal effects may have also played a role (e.g. Nakamura and Oke, 1988; Sini et al., 1996; Santamouris et al., 1999; Uehara et al., 2000; Bohnenstengel et al., 2004) especially at this sunlit part of the street. It is quite noticeable that low-speed winds in the canyon tend to blow North-eastwards. The spacing located near the station on the north side may have inuenced the wind direction, so that air owed between the two buildings. From Fig. 6a, we nd a linear relationship between the speed of winds in the canyon and that of winds above the roof. Winds above the roof move faster relative to those in the canyon. A linear regression line tted to the data for winds blowing from the E (70 to 125 ) and from the NW (270 to 335 ) had a coefcient of determination R2 0.62. This means that approximately 62% of the variation in the speed of winds, from E and NW, in the canyon is accounted for by the movement of winds at roof level. Similarly, considering winds from other directions, the coefcient of determination was R2 0.49. The inference made here is that a greater proportion of the variation in the speed of winds from other directions in the canyon cannot be explained by the dynamics of the winds above the roof. It was also observed (Fig. 6a) that winds from the east (E) were faster both in the canyon as well as at the roof level. However, from our study we could not conrm the simple linear relationship suggested by Nakamura and Oke

230

F. Ali-Toudert and H. Mayer

(1988) between the wind speed in the canyon and at roof level. Some studies suggest the existence of a threshold above which a coupling between the wind outside and inside the canyon may take place. In this study, a wind speed of about 5 ms1 (measured at 61 m) may be considered as a threshold, as shown in Fig. 6b, above which a strong correlation is found between the internal and external wind direction, whereas much more scattering is observed below this limit. By invoking the power law of wind prole, the corresponding wind speed directly above roof level (at 13 m a.g.l.) could be approximated to 2 ms1 , which agrees with estimates from previous studies (e.g. de Paul and Shieh, 1986; Nakamura and Oke, 1988).

4.2 Comfort analysis 4.2.1 Short-wave radiation ux densities The aspect ratio (H=W 1) and the street orientation (EW) together with the day of year and latitude are responsible for the solar exposure patterns prevailing within the canyon at street level. The north wall is almost permanently irradiated along with approximately one half of the canyon oor. In contrast, the south wall and the remaining half of the street surface are mostly shaded, except during early morning and late afternoon where they are irradiated as the sun crosses over the street. These patterns help to understand the following results relating to human heat gain within the canyon.

Fig. 7. Hourly mean values of short-wave radiation ux densities K (a) and (b) long-wave radiation ux densities L (b) received from the six directions surrounding a standing person located at the northern sidewalk of an EW oriented street canyon (H=W 1) as well as downgoing short-wave radiation at the roof level (53 m a.g.l.) on two typical summer days (14=15 July 2003) in Freiburg, Germany

Thermal comfort in an eastwest oriented street canyon under hot summer conditions

231

Figure 7a illustrates the short-wave radiation uxes (K) from the six directions on a person standing near the north wall. The importance of the orientation and the location within the canyon is evident. The downward irradiation (K #) recorded in the canyon has a daily course comparable to a location with a free horizon, i.e. the background station (K #roof). The only exceptions are before 0900 LST and after 1700 LST where the fac ades obstruct the direct solar beam. In the morning, the solar radiation comes from the south-east quadrant. As a result, K # is already at 820 Wm2 by 1100 LST and the radiation uxes coming from the east K (E !) and from the south K (S !) are also relatively high, about 460 Wm2 and 340 Wm2 , respectively (on 14 July). The maximum values of K # and K (S !) are recorded between 1200 and 1300 LST as the sun reaches its highest position (63 ) with the sun rays originating exactly from the south. The radiation uxes (K (W !) and K (E !)) measured parallel to the street axis are symmetrically reversed as the sun moves from the south-east quadrant to the south-west quadrant. Explicitly, K (W !) and K (E !) undergo the same pattern in the morning and afternoon, respectively. A sharp increase in the energy received from both directions is then recorded as the suns rays become parallel to the street axis, namely around 0900 LST for K (E !) and around 1700 LST for K (W !) with a maximum value of 660 Wm2 . In contrast, values not exceeding 130 Wm2 are measured before 1300 LST for K (W !) and after 1300 LST for K (E !). These correspond to diffuse and diffusely reected solar radiation components. Similarly, the short-wave radiation upwards K " as well as from the sunlit wall K (N !) do not exceed 125 W m2 and correspond to reected irradiation from the street and wall, (mean albedos of 0.15 and 0.13, respectively). 4.2.2 Long-wave radiation ux densities Solar exposure patterns inuence the long-wave radiation uxes (L) as shown in Fig. 7b. The asphalt road is mostly irradiated during the daytime and constitutes the highest source of longwave irradiance within the canyon at street level (L "). The peak value occurs between 1400 and 1500 LST and reaches 630 Wm2 , whereas the

lowest value is recorded at 0600 LST and equals 440 Wm2 . The radiant heat from the south facing wall L (N !) is also substantial as it is mostly irradiated. L (N !) shows a comparable temporal evolution as L ", though of lower magnitude. Indeed, being a vertical surface, it receives less short-wave radiation than the horizontal ground surface. Moreover, the asphalt pavement heats much more than the light coloured brick walls. After 1700 LST, the rate of heat release is slowed down as the canyon surfaces become shaded and thus cooler as less short-wave radiation is absorbed. The radiation ux densities from east L (E !) and west L (W !) are composed of heat released from the ground surface, the walls and the atmosphere. Yet, the inuence of the north wall has been dominant as the measurement was conducted close to it. In the morning, the amount of radiant heat from both directions is almost equal to that emitted by the north wall until 1300 LST and shows approximately the same increasing trend. However, the east side releases slightly more heat in the early afternoon while for the west side this occurs during the late afternoon, due to the sun exposure patterns previously mentioned. Because of the location of the radiation sensor close to the north wall, the heat emitted by the atmosphere, together with the street surface and the south wall, constitute a large part of L (S !). At 1100 LST, L (S !) is clearly lower than the other long-wave radiation ux densities. The main reason for this is the relatively low longwave atmospheric radiation L # combined with the small amount of radiant heat released by the opposite wall and part of the street still in shade. A rapid increase of L (S !) occurs in the afternoon. The peak value of about 580 Wm2 was recorded at 1500 LST as a result of the accumulated heat stored in the ground together with the late additional exposure of the whole street surface and south fac ade. The long-wave irradiance decreased from 1700 LST when all canyon facets became shaded. The slight increase of L (S !) at 0800 LST can be explained by the short exposure time of the southern side of the street in the early morning. At night, the street surface and the north wall remain the main sources of heat and show an

232

F. Ali-Toudert and H. Mayer

almost equal nocturnal cooling rate. The inuence of these two surfaces are also perceptible in the east L (E !) and west L (W !) uxes, and to a lesser extent in the south direction L (S !). 4.2.3 Radiant heat gained by a standing person In order to better understand the impact of the radiative environment described above on a human body, the actual short-wave and longwave radiation ux densities absorbed by a standing person in each direction have to be analysed. These ux densities take into account the human absorption coefcients and the human shape as expressed by Eq. (2). The totally absorbed short-wave radiation Kabs,total (Fig. 8a) ranges from 0900 to 1700 LST

between 160 and 200 Wm2 . On 14 July 2005, the highest values were recorded at 1100 LST (190 Wm2 ) and 1500 LST (200 Wm2 ) when the sun irradiated the human body laterally from the south-east or south-west directions. Because of the aspect ratio H=W 1 and the EW orientation of the street, the direct solar radiation before and after these two time points is blocked. This explains the sharp increase or decrease of energy gained at both times, respectively. At midday, i.e. between 1200 and 1300 LST, Kabs,total is somewhat lower due to the smaller body surface irradiated as the sun position is high. Basically, the human body absorbs less than 18 Wm2 from the wall Kabs (N !) which corresponds to diffusely reected solar radiation, whereas up to a maximum of 68 Wm2 is ab-

Fig. 8. Hourly mean values of short-wave radiation Kabs (a) and long-wave radiation Labs (b) absorbed by a standing person at the northern sidewalk of an EW oriented street canyon (H=W 1) on two typical summer days (14=15 July 2003) in Freiburg, Germany

Thermal comfort in an eastwest oriented street canyon under hot summer conditions

233

sorbed from the opposite side Kabs (S !) facing the sun in the early afternoon. The absorbed short-wave radiation from the south (Kabs (S !)) around noon, from the east (Kabs (E !)) in the morning, and from the west (Kabs (W !)) in the afternoon contributed substantially to Kabs,total, whereas the portion of the absorbed downward radiation (Kabs #) was distinctly lower than for Kabs (S !). Similar patterns were found for Kabs (E !) and Kabs (W !) but with the situation prevailing in the morning at the east side being symmetrically reected on the west side in the afternoon. The highest values of Kabs (E !) and Kabs (W !) reached 100 Wm2 at 0900 LST for the east and 1700 LST for the west facing side. With peak values about 5 Wm2 , the absorbed upward short-wave radiation (Kabs ") represents the lowest part of Kabs,total. In contrast to the absorbed short-wave radiation ux densities, the absorbed long-wave radiation ux densities (Fig. 8b) showed, as expected, a smoother daily evolution. Due to the cylinderlike shape of a standing person, the absorbed long-wave radiation coming from the lateral

directions was much higher than those directed upwards (Labs ") and downwards (Labs #). These vary from 85 Wm2 to 125 Wm2 for the lateral directions as opposed to between 20 Wm2 to 37 Wm2 in the vertical direction. Remarkably, the differences between Labs (E !), Labs (W !), Labs (N !) and Labs (S !) do not exceed 15 Wm2 . This means that the radiant environment is relatively homogenous vertically, in spite of the complex and variable shade patterns within the canyon. This is due to the fact that the radiant heat received from each direction originates from all surfaces (walls and ground) and from the sky simultaneously. The largest contrast however, is observed for Labs (S !) which shows the lowest amount (88 Wm2 ) at 1000 LST as the associated surfaces are still cool, and the highest value (124 Wm2 ) at 1500 LST, when the canyon surfaces have in the meantime the stored heat and receive additional energy from the irradiated opposite part of the street canyon. Altogether, the totally absorbed long-wave radiation Labs,total reaches distinctly higher values than Kabs,total, which emphasises the importance of the absorbed long-wave radiation for

Table 2. Percentage of short-wave radiation (SW) and long-wave radiation (LW) absorbed by a standing person at the south facing side of an EW oriented street canyon with an aspect ratio H=W 1 on two typical summer days (14=15 July 2003) in Freiburg, Germany LST (hrs) SW (%) LW (%) 11 29 71 12 16 74 13 24 76 14 26 74 15 27 73 16 27 74 17 24 76 18 3 97 19 2 98 20 1 99 21 to 5 0 100 6 2 98 7 4 96 8 16 84 9 28 72 10 29 71

Fig. 9. Relationship between the long-wave radiation absorbed by a standing person (Labs,total) at the northern sidewalk of an EW oriented street canyon (H=W 1) and the radiant heat emitted by the ground (L ") and the adjacent south facing wall L (N !) on two typical summer days (14=15 July 2003) in Freiburg, Germany

234

F. Ali-Toudert and H. Mayer

the radiant heat gained by a standing person. Labs,total peaked between 1400 and 1500 LST (540 Wm2 on 15 July 2005) and reached its minimum value between 0500 and 0600 LST (405 Wm2 ). At night, Labs,total provided all the radiant heat for a standing person within the urban street canyon, whereas during daylight hours (Kabs,total Labs,total) consisted of 74% Labs,total and 26% Kabs,total on average (Table 2). This highlights the importance of shading to reduce the radiant heat gained by a standing person, because it prevents the direct exposure of the person and keeps the adjacent surfaces cooler. The radiant heat gained by a standing person Labs,total is plotted against the heat released by the ground (L ") and the north wall (L (N !)) in Fig. 9. A strong linear relationship is found. The correlation of Labs,total was slightly higher with L (N !) than with (L ") because the street station was close to the north wall and because the body was considered in a standing posture (vertical). The importance of the ground has already been reported by Watson and Johnson (1988) and is probably more relevant for design purposes if these ndings could be extended to other locations across the street (i.e., further from). 4.2.4 Human thermal sensation The mean radiant temperature Tmrt, which sums up the absorbed short-wave and long-wave radiation uxes, is plotted in Fig. 10 together

with the comfort index PET and the air temperature Ta . A large daily amplitude of Tmrt was found, with a maximum of 66  C occurring at 1500 LST and a minimum of 20  C at 0500 LST on both days. The course of Tmrt can be easily understood in light of the two previous graphs (Fig. 8a and 8b). The exposure of a standing person to the short-wave irradiance is high from 0900 to 1700 LST, while long-wave irradiance becomes progressively high throughout the day, with a maximum occurring at 1500 LST. After 1700 LST, Tmrt decreases drastically because short-wave irradiance becomes negligible and results in a reduction of the surface temperatures and hence less radiant heat. At night, Tmrt values remain high, namely between 20  C and 30  C. This is attributable to the surplus heat released by the surfaces. During the day, Tmrt exceeded Ta to a large extent (Fig. 10). The peak value of the difference Tmrt Ta (34  C) occurred between 1400 and 1500 LST on 14 July 2004. As is known from other investigations, Tmrt is almost equal to Ta at night (e.g. Jendritzky et al., 1990; Mayer, 1993). During the day, the variation of the thermal index PET was mostly inuenced by Tmrt rather than by Ta (Fig. 10) although peak air temperatures and low-speed winds accentuated the PET values. This conrms that Ta alone is an inappropriate indicator for the assessment of thermal comfort outdoors during high pressure weather. A maximum PET value of 48  C was registered at 1600 LST on both days, while a minimum value of about 15  C was recorded.

Fig. 10. Hourly mean values of air temperature Ta, mean radiant temperature Tmrt and physiologically equivalent temperature PET on two typical summer days (14=15 July 2003) at the northern sidewalk of an EW oriented street canyon (H=W 1) in Freiburg, Germany

Thermal comfort in an eastwest oriented street canyon under hot summer conditions

235

5. Discussion and conclusion In this paper, the ndings of an in-situ study conducted in an eastwest oriented urban canyon, with an aspect ratio H=W 1, located in the mid-latitude city of Freiburg (Germany) in summer 2003, are presented. The near-surface microclimate within the urban street canyon and its impact on the thermal comfort of a standing person outdoors were investigated. The results are representative of typical hot and cloudless summer days in a mid-latitude location. Measurements were taken of the meteorological variables required for determining the thermal index PET, used here as a thermophysiologically signicant index. The evolution of the air temperature Ta observed within the canyon was in good agreement with previous eld studies. The air temperature Ta was found to be slightly affected by the canyon geometry except close to irradiated canyon facets where the air was warmer. A sensitivity analysis carried out on the basis of the obtained data also revealed that PET increases at a rate of about 0.75 K per unit increase in Ta (PET 3 Ta), if all other parameters were 4 kept constant. It is, therefore, inappropriate to use Ta as the main indicator for comfort outdoors under hot and sunny conditions. This contrasts with the common use of Ta as a comfort indicator (e.g. Swaid et al., 1993; Shashua-Bar and Hoffamn, 2000; Coronel and Alvarez, 2001; Grundstrm et al., 2003). Despite the small o data sample, the wind ow is found to be strongly correlated in speed and direction with the free air and corroborates former eld studies. The results showed the dominant effect of the exposure to the sun on the human thermal discomfort in the daytime with Tmrt and PET attaining their maxima around 66  C and 48  C, respectively. This was mainly due to the large amounts of energy absorbed by an irradiated standing person, namely up to a total of 730 Wm2 . On the northern side of the investigated EW canyon, with H=W 1 and typical urban materials, the pedestrian absorbed about 74% of the total energy as long-wave irradiance (405 to 545 Wm2 ) compared to 26% as short-wave irradiance (160 to 200 Wm2 ) in the daytime. The absorption of energy from the sun and the surroundings depends strongly on the aspect ratio and orientation and

conrms former numerical results (Ali-Toudert and Mayer, 2006). Further eld investigations are required in order to verify the generality of these results for other locations and climatic conditions. A recent one, for instance, was conducted in a subtropical desert city (Ali-Toudert et al., 2005). Based on a similar methodology for assessing outdoor thermal comfort, this study corroborated many of the ndings of the present study: the decisive role of the aspect ratio and orientation on the resulting thermal sensation, the conservative character of Ta, the proportions of shortwave and long-wave irradiances absorbed by a human body and the effects of building materials. This can partly be explained by the extreme hot conditions that prevailed at the time of study in Freiburg, which is rather typical for lower latitudes. Intuitively, this study quantitatively conrms that shading for both pedestrians and surrounding surfaces is crucial in mitigating human heat stress. A judicious combination of high aspect ratios and orientation, arranging galleries, planting trees, greening the fac ades or using other shading devices on the walls are a few possible solutions. Promoting ventilation through appropriate urban plan orientation and density is a further possibility to reduce heat stress. These were veried quantitatively to be effective for the subtropics (Ali-Toudert, 2005). Short-wave irradiance absorbed by a pedestrian is responsible for peak discomfort values. Its impact can be reduced by wearing light coloured clothes (ak <0.7). Moreover, special attention has to be given to the surfaces themselves: ground pavements constructed of light coloured, porous materials are advisable (Ca et al., 1998); thin layer pavements or pavements mixed with green surfaces for promoting evaporation from underground are also suitable (Asaeda and Ca, 1993; Aseada et al., 1996) especially in latitudes where summers are not dry, like in Freiburg. Building materials also play a role: high thermal capacity may help to reduce surface temperatures further and thus the heat released. However, this could lead to the delayed cooling of deep street canyons (Ali-Toudert et al., 2005). Finally, as the surrounding radiative environment is decisive in determining outdoor human

236

F. Ali-Toudert and H. Mayer Ca VT, Aseada T, Ashie Y (1998) Utilization of porous pavement for the improvement of summer urban climate. u Proc. 2nd Japanese-German meeting. Klimaanalyse fr die Stadtplanung. Rep Research Centre for Urban Safety and Security. Kobe University, Sp. Rep. No. 1: 196177 Coronel JF, Alvarez S (2001) Experimental work and analysis of conned urban spaces. Solar Energy 70: 263273 Eliasson I (1993) Urban climate related to street geometry. PhD thesis. University of Gothenburg. Dept Phy Geogr GUNI rapport 33 Ernst S (1995) Tagesperiodische Windsysteme und Beluftungsverhltnisse in Freiburg i. Br.-Planungsrelevante a Aspekte eines Bergwindsystems. Freiburger Geographische Hefte 49 (in German) Fink A, Brcher T, Krger A, Leckebusch GC, Pinto JG, u u Ulbrich U (2004) The 2003 European summer heatwaves and drought synoptic diagnosis and impacts. Weather 59: 209216 Grundstrm K, Johansson E, Mraisi M, Ouahrani D (2003) o Climat et urbanisme la relation entre confort thermique et la forme du cadre b^ti. Report 8. Housing Development a and Management. Lund University. Sweden Hppe P (1992) Ein neues Verfahren zur Bestimmung der o mittleren Strahlungstemperatur im Freien. Wetter und Leben 44: 147151 Hppe P (1993) Heat balance modelling. Experientia 49: o 741746 Hppe P (1999) The physiological equivalent temperao ture a universal index for the biometeorological assessment of the thermal environment. Int J Biometeorol 43: 7175 Hussain M, Lee BE (1980) An investigation of wind forces on the 3D roughness elements in a simulated atmospheric boundary layer ow. Part II Flow over large arrays of identical roughness elements and the effect frontal and side aspect ratio variations. Department of Building Science. Univ. of Shefeld, UK Jendritzky G, Menz G, Schirmer H, Schmidt-Kessen W (1990) Methodik zur rumlichen Bewertung der thera mischen Komponente im Bioklima des Menschen. Fortgeschriebenes Klima-Michel-Modell. Beitrge der a Akademie fr Raumforschung und Landesplanung. u Hannover, Vol. 114 Kim J-J, Baik J-J (1999) A numerical study of thermal effects on ow and pollutant dispersion in urban street canyons. J Appl Meteorol 38: 12491261 Mayer H (1993) Urban bioclimatology. Experientia 49: 957963 Mayer H, Hppe P (1987) Thermal comfort of man in o different urban environments. Theor Appl Climatol 38: 4349 MIF (2005) Website of the Meteorological Institute Freiburg, urban climate station, http:==www.mif.unifreiburg.de=station Nagara K, Shimoda Y, Mizuno M (1996) Evaluation of the thermal environment in an outdoor pedestrian space. Atmos Environ 30: 497505 Nakamura Y, Oke T (1988) Wind, temperature and stability conditions in an eastwest oriented urban canyon. Atmos Environ 22: 26912700

thermal comfort, special emphasis was placed here on an extensive measurement of the shortwave and long-wave radiation ux densities within the urban street. Given the lack of such information, the data gathered in this study could be used for validation purposes, such as the numerical estimation of Tmrt, e.g. by the model of Asawa et al. (2004) and ENVI-met model (Bruse, 2005).
Acknowledgements The authors are grateful to Dr. Thomas Holst, Dr. Jutta Rost and Dr. Florian Imbery for helping in planning and carrying out the measurements. The text was proofread by Dr. Argwings Ranyimbo. References Ali-Toudert F (2005) Dependence of outdoor thermal comfort on street design in hot and dry climate. Rep Meteor Inst Univ Freiburg No. 15, http:==www.freidok.unifreiburg.de=volltexte=2078 Ali-Toudert F, Mayer H (2006) Numerical study on the effects of aspect ratio and solar orientation on outdoor thermal comfort in hot and dry climate. Building and Environment 41: 94108 Ali-Toudert F, Djenane M, Bensalem R, Mayer H (2005) Outdoor thermal comfort in the old desert city of BeniIsguen, Algeria. Climate Research 28: 243256 Arneld J (1990) Street design and urban canyon solar access. Energy and Buildings 14: 117131 Arneld J (2003) Two decades of urban climate research: a review of turbulence, exchanges of energy and water, and the urban heat island. Int J Climatol 23: 126 Arneld J, Mills G (1994) An analysis of the circulation characteristics and energy budget of a dry, asymmetric, eastwest urban canyon. II. Energy budget. Int J Climatol 14: 239261 Asaeda T, Ca VT (1993) The subsurface transport of heat and moisture and its effects on the environment: a numerical model. Bound-Layer Meteor 65: 159179 Asaeda T, Ca VT, Wake A (1996) Heat storage of pavement and its effect on the lower atmosphere. Atmos Environ 30: 413427 Asawa T, Hoyano A, Nakaohkubo K (2004) Thermal design tool for outdoor space based on numerical simulation system using 3D-CAD. Proc. 21th Conf. on PLEA, Eindhoven. Netherlands: 10131018 ASHRAE (2001) Chapter 13 Measurements and instruments. In: Handbook of Fundamentals. American Society for heating Refrigerating and Air Conditioning. Atlanta: 13.2613.27 Bohnenstengel S, Schlnzen KH, Graw D (2004) Inuence u of thermal effects on street canyon circulations. Meteorologische Zeitschrift 13: 381386 Bruse M (2005) ENVI-met website. http:==www. envi-met.com

Thermal comfort in an eastwest oriented street canyon under hot summer conditions Nikolopoulou M, Lykoudis S (2005) Thermal comfort in outdoor urban spaces: analysis across different European countries. Building and Environment (in press) Nbler W (1979) Konguration und Genese der Wrmeinsel u a der Stadt Freiburg. Freiburger Geographische Hefte. Heft 16 (in German) Nunez M, Oke TR (1977) The energy balance of an urban canyon. J Appl Meteor 16: 1119 Oke T (1981) Canyon geometry and the nocturnal urban heat island: comparison of scale model and eld observation. J Climatol 1: 237254 Oke T (1988) Street design and urban canopy layer climate. Energy and Buildings 11: 103113 de Paul FT, Shieh CM (1986) Measurements of wind velocity in a street canyon. Atmos Environ 20: 455459 Pearlmutter D, Bitan A, Berliner P (1999) Microclimatic analysis of compact urban canyons in an arid zone. Atmos Environ 33: 41434150 Pickup J, de Dear R (1999) An outdoor thermal comfort index (OUT-SET ) -Part 1 The model and its assumptions. Proc. 15th Int. Congr. Biometeorol. & Int. Conf. Urban Climatol, Sydney, Australia: 279283 Santamouris M, Papanikolaou N, Koronakis I, Livada I, Asimakopoulos D (1999) Thermal and air ow characteristics in a deep pedestrian canyon under hot weather conditions. Atmos Environ 33: 45034521 Shashua-Bar L, Hoffman ME (2000) Vegetation as a climatic component in the design of an urban street. Energy and Buildings 31: 221235 Sini J-F, Anquetin S, Mestayer P-G (1996) Pollutant dispersion and thermal effects in urban street canyon. Atmos Environ 30: 26592677 Spagnolo J, de Dear R (2003a) A eld study of thermal comfort in outdoor and semi-outdoor environments in subtropical Sydney Australia. Building and Envir 38: 721738 Spagnolo J, de Dear R (2003b) A human thermal climatology of subtropical Sydney. Int J Climatol 23: 13831395

237

Staiger H, Bucher K, Jendritzky G (1997) Perceived temperature. The thermophysiologically right assessment of heat and cold stress staying outdoors in the unit degree celsius. Ann Meteorol 18: 100107 (in German) Stathopoulos T, Wu H, Zacharias J (2004) Outdoor thermal comfort in an urban climate. Building and Environment 39: 297305 Svensson MK, Thorsson S, Lindqvist S (2003) A geographical information system model for creating bioclimatic maps examples from a high, mid-latitude city. Int J Biometeorol 47: 102112 Swaid H, Bar-El M, Hoffman ME (1993) A bioclimatic design methodology for urban outdoor spaces. Theor Appl Climatol 48: 4961 Todhunter PE (1990) Microclimatic variations attributable to urban canyon asymmetry and orientation. Phys Geogr 11: 131141 Uehara K, Murakami S, Oikawa S, Wakamatsu S (2000) Wind tunnel experiments on how thermal stratication affects ow in and above urban street canyons. Atmos Environ 34: 15531562 Watson ID, Johnsson GT (1988) Estimating person viewfactors from sh eye lens photographs. Int J Biometeorol 32: 123128 Wedding JB, Lombardi DJ, Cermak JE (1977) A wind tunnel study of gaseous pollutants in city street canyons. J Air Pollut Contr Ass 27: 557566 Yoshida A, Tominaga K, Watani S (1990=91) Field measurements on energy balance of an urban canyon in the summer season. Energy and Buildings 15=16: 417423

Authors address: Fazia Ali-Toudert (e-mail: fazia. alitoudert@meteo.uni-freiburg.de), Helmut Mayer, Meteorological Institute, University of Freiburg, Werderring 10, 79085 Freiburg, Germany.

You might also like