Next Article in Journal
A Quantitative Method to Measure and Speciate Amines in Ambient Aerosol Samples
Next Article in Special Issue
Changes Detected in Five Bioclimatic Indices in Large Romanian Cities over the Period 1961–2016
Previous Article in Journal
Impact of Land Cover and Leaf Area Index on BVOC Emissions over the Korean Peninsula
Previous Article in Special Issue
UV-Index Climatology for Europe Based on Satellite Data
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Accuracy of Mean Radiant Temperature Derived from Active and Passive Radiometry

by
Henning Staiger
1,* and
Andreas Matzarakis
2
1
Emeritus Deutscher Wetterdienst, Corneliusstr. 2a, 64546 Mörfelden-Walldorf, Germany
2
Research Centre Human Biometeorology, Deutscher Wetterdienst, Stefan-Meier-Str. 4, 79104 Freiburg, Germany
*
Author to whom correspondence should be addressed.
Atmosphere 2020, 11(8), 805; https://doi.org/10.3390/atmos11080805
Submission received: 9 June 2020 / Revised: 24 July 2020 / Accepted: 27 July 2020 / Published: 30 July 2020
(This article belongs to the Special Issue Challenges in Applied Human Biometeorology)

Abstract

:
The concept of the mean radiant temperature (Tmrt) allows the study of radiative exchanges between a human and its environment. It presupposes that the radiant effects on the person of the actual environment, which is generally heterogeneous, and the virtual environment, which is defined as homogeneous, are identical. ISO 7726 specifies the required accuracy in Tmrt as input of rational thermal indices, outdoors ±5 (K). Tmrt accounts for the radiant heat absorbed by skin/clothing from the shortwave (SW) and longwave (LW) spectral bands. Most of the radiant components are isotropic. However, there are anisotropic SW components; namely the direct irradiance and under clear or partly obstructed skies a significant circumsolar fraction (fcs) in the diffuse irradiance. Both originate from the close proximity of the solar disk. This study highlights the effect of fcs on Tmrt. In the scope of human biometeorology a standing body posture is standard. For unidirectional irradiances its radiant cross-section varies dependent on the solar altitude. Active radiometry in deriving Tmrt is based on measured irradiances. One method is the Klima-Michel-Modell (KMM) that uses readily available measurements from standard meteorologically radiant observations. KMM references Fanger’s area projection factors that are derived from precise measurements of real humans. Thus, KMM serves as reference in evaluation of further methods. One is the six-directional instrument (Tmrt,r,6−Dir). Slightly simplifying a standing human, it represents a subject as a rectangular solid. Tmrt,r,6−Dir is derived based on measured irradiances incident on the vertical and horizontal planes. In passive radiometry the energy balance equation of a black globe thermometer is solved that leads to Tmrt,Tg,BG. fcs significantly impacts Tmrt with noticeably reduced values for high and increased for low solar altitudes. Hence, accounting for fcs is essential for the accuracy of Tmrt. For KMM an extension to an existing algorithm is provided in order to include fcs into the Tmrt calculation that results in Tmrt,r,KMM. For Tmrt,r,6−Dir the radiant cross-section of the solid depends to a minor extent on its azimuth relative to the solar azimuth. As a result Tmrt,r,6−Dir slightly scatters compared to Tmrt,r,KMM. However, it remains within ±2 (K). Tmrt,Tg,BG compared to Tmrt,r,KMM complies only at night with the ISO 7726 bin of ±5 K. Tmrt,Tg,BG significantly overestimates Tmrt,r,KMM during the daytime, because of its greater SW absorptance compared to skin/clothing and to a smaller extent because the standing posture is represented by a sphere. Particularly in sunny conditions, Tmrt,Tg,BG is subject to considerable variance. Thus, outdoors during the daytime, Tmrt,Tg,BG is unable to serve as an appropriate input for the calculation of rational-based thermal indices.

1. Introduction

Human biometeorology is the interdisciplinary field of science that studies interactions between the human body and the Earth’s atmosphere and it is thus related to outdoor environments. Especially, it is concerned with contingency planning, health protection, regional and urban planning, the design of open spaces, various aspects of tourism and recreation areas, as well as research in climate change [1]. Its results are used in heat warnings and to provide medical-related information for weather forecasting. All these facets require a rational, thermophysiologically consistent assessment of the human thermal environment. This is frequently accomplished by applying thermal indices. Rational thermal indices are based on complete human energy budget models consisting of both a controlled passive and a controlling active system [2]. They require six basic parameters in order to define the human thermal environment. These are four meteorological variables, namely air temperature (ta (°C) or Ta (K)), water vapor pressure (ea (hPa)), air velocity (va (m/s)) and mean radiant temperature (Tmrt (K) or tmrt (°C)), plus two inputs that describe properties affecting the human body: the insulation of clothing worn and the activity of the person described in terms of net metabolic heat production [3].
Compared to ta, ea and va, which can readily and accurately be measured, Tmrt is the most complex of the four meteorological inputs [1,4]: it is the uniform temperature of an imaginary black enclosure (emissivity ε = 1), in which a human body of defined azimuth and posture would gain the same radiant heat as in the actual nonuniform enclosure [5,6,7,8]. Tmrt assesses the impact of the solar shortwave (SW, 0.3–3 µm) and the terrestrial (thermal) longwave (LW, 3–50 µm) radiant components on the human energy balance.
The concept of Tmrt allows the study of radiative exchanges between a human person and its environment. Because it presupposes that the radiant effects on the person of the actual environment which is generally heterogeneous and the virtual environment which is defined as homogeneous are identical, it facilitates thermophysiological modeling [5,8]. Tmrt is a direct measure of the environmental sensible heat stress onto the body that come from radiation. The rational indices resulting from thermophysiological modeling and their application require sufficient accuracy of Tmrt. Those are quantified by ISO 7726 according to the extent of the thermal annoyance to be assessed.
The calculation of Tmrt is originally developed for indoor applications. In a room, Tmrt may be calculated from the surface temperatures TN of the surrounding N surfaces with uniform temperature each by accounting for the angle factor FpN between a person and the surrounding surfaces [6,8,9]:
T mrt 4 = T 1 4 · F p 1 + T 2 4 · F p 2 + + T N 4 · F p N
FpN is a function of the shape, the size and the relative position of surface N in relation to the person. The position of the surface N is described in spherical coordinates by azimuth and elevation angles. The angle factor can precisely but elaborately be determined by numerical integration over the azimuth and elevation angles. The integration takes place between the normal of the person’s location on the concerned surface up to the angle of N. FpN consists of the projected area factor fp that is weighted according to the relative position of surface N to the person. fp = Ap/Aeff relates the projected area of a person Ap to its effective radiant area Aeff, i.e., the total surface of the person minus that area parts that overlap, e.g., arms and trunk. Thus, fp is a function of the body posture and the elevation angle [6,10,11] and is precisely determined by the shadow area of the irradiated person on a plane normal to the direction of the incident beam. As the sum of the angle factors is unity, the fourth power of Tmrt will be seen to be equal to the mean value of the surrounding surface temperatures to the fourth power, weighted with the respective angle factors [6].
By inversion of the Stefan–Boltzmann law, the T N 4 in Equation (1) can be replaced by measured irradiances, i.e., by active radiometry. Thus, measured radiant flux densities of the effective components in the SW and LW spectral bands can directly be applied. This is the base for outdoor use within the scope of human biometeorology where surface temperatures of the upper (atmospheric) half space are in principle not available.
In applied human biometeorology an upright reference subject is used that is rotationally symmetric to its vertical axis because the azimuthal orientation of the subject is usually unknown. Its elevation dependent fp is derivable from diagrams and tables [1,6,11]. Because only the fraction of radiant energy that is absorbed by the skin/clothing surface can have an effect on the human energy balance, the clothing/skin of the reference subject is assumed to show overall SW absorption (αsw,Sk) of 0.70 and LW emissivity (εlw,Sk) of 0.97 that is identical with its absorptivity (Kirchhoff law) [1,8,9].
In the scope of human biometeorology, two methods that apply active radiometry are in use. One [12] bases on operationally available radiant flux densities as they, for example, are recorded by meteorological services according to guidelines of the World Meteorological Organization [13,14]. The other applies the definition of plane radiant temperature [8,9]. It uses measured SW and LW radiant flux densities onto the six planes of a rectangular solid that represents a human of standing body posture [1,15].
Another approach to determine Tmrt is passive radiometry. Here, Tmrt is indirectly derived by determining the energy balance of a measuring system, especially the black globe thermometer [8,16,17]. The method has a long history and provides in those early years at least estimates of Tmrt where the measurement of radiant flux densities was complex due to technical reasons and expansive [16,17]. The black globe was originally introduced in order to measure radiant heat indoors in relation to human comfort [17]. However, it is also applied outdoors, where it is integrated in instruments that provide the measured base for the calculation of the wet-bulb-globe-temperature index [18,19]. In the scope of human bioclimatology, the black globe is now also being used outdoors in the form of a matt black-painted table-tennis ball [20,21]. Tmrt derived from the black globe is an approximation due to the difference in shape between a person and a globe [8]. In particular, direct SW irradiation at high solar altitudes will be over-estimated by the globe in relation to that received by a standing person. The black globe differs strongly in the SW absorptance compared to the outer surface of a clothed person. Furthermore, the accuracy of measuring Tmrt using a black globe can vary to a great extent according to the values for the other characteristics of the environment [8,22].
The occurrence of heat stress situations is usually related to clear sky conditions, when humans are directly exposed to intense, frequently asymmetric irradiation. In this case, the diffuse radiant intensity from the sun’s direction is substantially higher than from the remaining upper half space [23,24,25] and may significantly affect Tmrt especially for the standing body posture.
This study aims to assess the three methods for outdoor use. In particular, it shall be evaluated whether the achieved accuracy is sufficient that the derived Tmrt can serve as input for the calculation of thermal indices. Furthermore, it shall be investigated how the circumsolar anisotropy in the diffuse radiant component affects Tmrt.
The objectives are accomplished by defining the effective solar and terrestrial irradiances in section “Methods and Materials”, in which the methods and their theoretical background are shortly described and the available data for their evaluation are presented. The assessment methodology including the method selected for reference is introduced. Based on these data, the accuracy of the remaining methods is evaluated against the reference in section “Results” and the issue is examined how the circumsolar anisotropy in the diffuse SW irradiance affects Tmrt.

2. Methods and Materials

2.1. Effective Solar and Terrestrial Irradiances

Solar electromagnetic irradiation that reaches the earth’s surface covers a spectral range between ~280 and 3000 nm. There are different impacts on the human organism depending upon the spectrum. The wavelengths between 400 and 730 nm enable photopic and scotopic vision. Especially in the UV-B range ≤ 315 nm, the impact on the skin may be harmful. Dependent on the dose and the skin type, it can provoke an erythema. However, the predominant impact of SW and LW irradiance on the skin is thermal. It is spectrally independent as long as the intensities in a given spectral range are less than 1000 (W/m2). Thus, it is the total intensity or energy that affects a person’s thermal psycho-physiological state [26]. Hence, it is to be concluded that Tmrt can be derived for the relevant wavelength bands from spectrally integrating instruments. They are deployed as standard in meteorology.
The effective SW radiant flux densities (W/m2), are
  • The downward direct irradiance (EI) incident on a plane normal to the radiation’s direction and measured by a pyrheliometer. EI is that part of the solar radiation that is not scattered by atmospheric constituents in its transfer to the earth’s surface. Thus, EI is unidirectional because it originates from the solar disk, i.e., it is anisotropic.
  • The downward diffuse (directionally indistinguishable) irradiance (ED) incident on a horizontal plane facing the upper half space and measured by a pyranometer combined with e.g., a shadow disk moved by a sun tracker or a shadow band. The latter requires subsequent correction of the reading, which accounts in particular for sky radiance anisotropy [23,27,28]. ED is that part of the solar radiation that has been subjected to scattering by atmospheric atoms and molecules (Rayleigh) and by aerosol particles and cloud droplets (Mie). Scattering changes the direction of the incident beam, yet leaves its energy and frequency unchanged. ED is multidirectional; frequently it impacts with the same amount of energy from all directions of the sky dome, and is then termed isotropic.
  • The surface reflected irradiance (ER) above the lower half space and incident on a horizontal plane facing downward and measured by a pyranometer. ER is predominantly Lambertian equivalent reflected, i.e., it is diffuse. The direct mirror-like reflected proportion of ER remains small and can be regarded as insignificant [29,30]. Thus, ER is usually isotropic with regard to the lower half space. Exceptions to Lambertian reflection include water and ice surfaces.
However, the most frequently observed irradiance is the SW global irradiance (EG) incident on a horizontal plane facing upward (pyranometer). EG is defined as EG = EI∙sin(γs) + ED, where γs (degrees) is the solar elevation. Hence, measured EG and ED allow EI to be determined, the measurement of which is more difficult. There are algorithms available that break measured EG down into its fractions using a stochastic approach (see e.g., [31]).
ER can also be attributed to EG by ER = ρsEG. ρs is the surface albedo (0–1) that depends on ground surface properties. For a given location, ρs can frequently be regarded as constant, but there may be seasonal variations [32], e.g., due to snow cover in winter.
Under overcast skies, ED is completely isotropic. However, under skies which are clear or only with low cloud cover, ED includes a significant anisotropic fraction that is concentrated around the solar disk with a view angle of ≤12 (degree). This circumsolar brightening originates from Mie scattering caused by aerosols and cloud particles that has a narrow peak in the forward direction. This anisotropic circumsolar fraction (fcs) of ED is never available from operational observations. Hence, it must be derived using sky radiance models, e.g., that of Perez et al. [23], which accounts for γs, sky clearness (overcast to clear) and sky brightness (cloud thickness and/or aerosol loading ranging from low to high). fcs is especially important if the irradiance on tilted surfaces is calculated (see e.g., [32]) because this depends on the azimuth of the radiation source, e.g., for an upright human.
The amount of radiation which each body emits depends on its (surface) temperature. Hence, the atmosphere and the ground are sources of electromagnetic radiation due to their temperature in the thermal-spectral range. The effective LW flux densities (W/m2), measured by pyrgeometers, are
  • The isotropic diffuse downward thermal irradiance of the atmosphere (EA) incident on a horizontal plane facing the upper half space;
  • The upward isotropic diffuse irradiance of the lower half space (EE) incident on a horizontal plane facing the lower half space. EE includes the small fraction (≤5%) of ground-reflected EA.

2.2. Klima-Michel-Modell

The calculation of Tmrt inside the Klima-Michel-Modell (KMM) [12] is based on Equation (1) for outdoor use. Breaking the unidirectional irradiances originating from the area surrounding the solar disk down into a vertical and a horizontal component the respective angle factor FpN (Equation (1)) can be shown to equal fp(γs), because the weight in FpN that accounts for the relative position of surface N to the person is in this case unity. For the standing posture of a person rotationally symmetric to its vertical axis a parameterization for fp based on diagrams in [6] can be applied (Figure 1) [12]:
f p = 0.308 × cos { γ s × ( 0.998 γ s 2 50000 ) } .
For a surrounding free of significant horizon limitations [33], Tmrt can, therefore, be derived directly from measured radiant flux densities, active radiometry, referred to below by “r” in subscripts. Because the view angle of the circumsolar fraction of ED is small, it can reasonably be treated like EI as originating from the solar disk [24]. Then, the equation of Jendritzky et al. [12] for the calculation of Tmrt, which assumes fcs = 0, can be extended for all fcs. It reads:
T mrt , r , KMM = { 1 σ SB · [ f a · ( E A + E E ) + α sw , Sk ε lw , Sk · (   f a · ( E R + ( 1 f cs ) · E D ) + f p · ( E I + 1 sin ( γ s )   × f c s · E D ) ) ] } 0.25
σSB = 5.67∙10−8 (W∙m−2∙K−4) is the Stefan–Boltzmann constant. Because isotropic diffuse radiation from one of two half spaces irradiates the body with the same intensity over all azimuth and elevation angles (Equation (1)), the integration over all angle factor FpN of this half space is fa = 0.5, where fa is independent of γs and the shape of the irradiated body [12,24,34]. Hereafter, the procedure based on Equation (3) is called Tmrt,r,KMM, relating it to the Klima-Michel-Modell (KMM) [12].

2.3. Six-Direction Instrument

Another method for determining Tmrt using observed irradiances is the six-directional instrument (Tmrt,r,6-Dir (K)) that posits the subject concerned as a rectangular solid [1,15]. This Tmrt,r,6-Dir (K) is also defined as the plane radiant temperature [8,9], however using pyranometers and pyrgeometers facing up-down, left-right, forwards-backwards and belongs so to active radiometry. The six instruments of each type may be replaced by one that can be rotated in the six directions to deliver a temporal sequence of observations. The six weighting factors (Wi) for the observed SW (Ki) and LW (Li) radiant flux densities (W/m2) are derived using Fanger’s formula [6] for fp of a rotationally symmetric person who is standing. Wi equals 0.06 for up and down and 0.22 for the four cardinal points. The energy absorbed by the rectangular solid (Sr) is ∑6i=1 Wi∙(αsw,SKKi + εlw,SKLi) (W/m2) and, thus
T mrt , r , 6 Dir = ( S r σ SB · ε lw , Sk ) 0.25

2.4. Black Globe Thermometer

Tmrt can be derived indirectly by solving the energy balance equation of a measuring system that is affected by convection, conduction and radiation (passive radiometer). Tmrt is derivable when the system is in equilibrium, i.e., the radiant energy input is exactly balanced by convective heat dissipation [16]. The black globe thermometer is most frequently used for this [17]. This is essentially a hollow spherical globe containing an ordinary thermometer. The bulb of this thermometer is at the center of the sphere and theoretically represents the mean surface temperature of the sphere’s inner surface. This integral value is composed of individual values of SW and LW radiant flux densities of the actual environment that vary with the solid angle. However, they may be smoothed, dependent on the heat conductivity of the globe’s material.
The usual form of the black globe thermometer (BG) consists of a thin-walled (≈0.4 mm) copper sphere of diameter (dg) of 0.15 (m) that is coated in matt black paint, emissivity εg = 0.95 [8,22]. When exposed to radiation, such black globe thermometers usually register a higher temperature, the globe temperature (Tg (K) or tg (°C)), than an ordinary, radiation-protected thermometer that results in ta. Generally, the transient heat balance of a globe at time t is [35,36]:
( Q r ( t ) Q c ( t ) ) outer ( Q r ( t ) Q c ( t ) ) inner = A g · δ g · ρ · c · d T g d t
Qr(t) and Qc(t) are the radiant and the convective heat transfer rates, respectively, evaluated for the outer and for the inner surface of the globe. The sphere’s surface (Ag) is 4∙π∙dg2 (m2), and δg (m) is the shell’s thickness. ρ (kg/m3) is the mass density of the shell’s material, and c (J K−1 kg−1) its specific heat capacity. dt is a time interval. This balance equation neglects heat conduction across the surface. This is only justified for metallic (high thermal diffusivity) shells that are thin relative to their overall dimensions [35]. Over the whole globe, the heat loss from the inner surface is zero because elements of the inner globe radiate out to each other and internal, free convection air currents redistribute heat from one point to another [35,37]. However, the radiant properties of the inner surface and the mass of the thermometer’s bulb affect the time constant of the instrument. The energy balance on the globe’s outer surface yields [8,22,35,38,39]:
( Q r ( t ) Q c ( t ) ) outer = σ SB · ε g · ( T mrt 4 T g 4 ) h ¯ c , g · ( T a T g ) = δ g · ρ · c · d T g d t
h ¯ c , g (W∙m−2∙K−1) is the convective heat transfer coefficient averaged over the surface of the sphere. In equilibrium, the right hand side of Equation (6) equals zero, lim t ( Q r ( t ) Q c ( t ) ) outer = 0, and, then, Equation (6) solves
T mrt , Tg = ( T g 4 + h ¯ c , g σ SB · ε g · ( T g T a ) ) 0.25
Outdoor forced convection dominates for va ≥ 0.03∙(TgTa)0.42 [38]. According to the similitude theory h ¯ c , g = κ · N u ¯ / L   . N u ¯   (dimensionless) is the mean Nusselt number, i.e., the ratio of mean convective h ¯ c , g to conductive heat transfer κ (W∙m−2∙K−1) in the film layer, here at the boundary between the air and the globe’s shell. L (m) is the characteristic length, for the sphere it is dg. N u ¯ is additionally a function of va on the sphere and of air properties, namely kinematic viscosity, specific heat capacity, thermal conductivity and density that are usually summarized in the value 6.3 [8,22] for standard atmospheric conditions. Thus, the black globe equation for forced convection (6) reads
T mrt , Tg , BG ( T g 4 + 6.3 σ SB · ε g · v a 0.6 d g 0.4 · ( T g T a ) ) 0.25

2.5. Data

In order to realistically compare Tmrt,r,6-Dir against Tmrt,r,KMM, observations are required that cover in particular a sufficient number of solar altitudes and weather conditions. Such large amounts of observations are not available under all weather conditions for the 6-Dir instrument. Therefore, the fact is exploited that direct, diffuse and surface reflected SW and thermal LW irradiances onto tilted surfaces can readily be calculated based on (solar) geometry from standard meteorological irradiance measurements (see e.g., [32]). The sole exception in the transposition to tilted surfaces is the anisotropic fraction in ED. No such measurement is available and, thus, has to be modeled. To this end, a three component treatment of sky diffuse irradiance is applied based on known values of ED, EG and γs. [23,25]. The circumsolar fraction fcs is then treated like EI, i.e., dependent on solar geometry. The small band of horizon brightness is added to the isotropic fraction in ED. ER is regarded as exclusively isotropic diffuse, like EA and EE. Hence, for an environment with no significant horizon limitations [33], the SW and LW flux densities on the vertical surfaces of the 6-Dir instrument can be derived from standard horizontal and beam irradiance measurements based on more or less exclusively geometric relations.
Tmrt is applied as input for the calculation of thermal indices that represent steady state thermophysiological values of a reference person. Because steady state is usually achieved within 2 h [7,40], hourly sums of irradiation are the appropriate radiant input to characterize the radiant conditions in the adaptation phase leading to steady state. Table 1 lists the radiant components and auxiliary quantities that are available in hourly resolution for the years 2006 and 2007 from WMO Baseline Solar Radiation Network (BSRN) sites [13]: Ny Alesund (NYA), Norway: 78.93° N, 11.95° E, 11 m MSL; Lerwick (LER), United Kingdom: 60.13° N, 1.18° W, 84 m MSL; Payerne (PAY), Switzerland: 46.82° N, 6.94° E, 491 m MSL, Sde Boqer (SBO), Israel: 30.91° N, 34.78° E, 500 m MSL. BSRN data offer the highest available accuracy [41]. All BSRN sites fulfill the conditions that they are free from immediate local effects such as pollution, heat island effects, local orography such as surface tilt and elevate horizon, and isolated atmospheric conditions such as local cloud developments [13].
The missing ER for LER and SBO is replaced by assuming ρs = 0.2. The missing EE for both sites is modeled using a Bowen approach in solving the energy balance at the earth’s surface [12]. The latter requires auxiliary quantities from synoptic meteorological observations (SYNOP) that are only available in very limited quantities for SBO. Because these are limited to 06 and 18 UTC, radiant components for high solar altitudes (noon) cannot be modeled.
For the years 2001 to 2003, hourly radiant components for Rostock-Warnemuende (RWA), Germany: 54.18° N, 12.08° E, 4 m MSL are also available (Table 1). Their accuracy is in line with long-term measurements of national meteorological services [41]. The measuring field is located in the dunes bordering the Baltic Sea. Here, the SW ER is measured accounting for the higher albedo of the dune sand and the LW components EA and EE are measured based on Schulze pyrradiometers. Compared to pyrgeometers, the measurements should show approximately identical means, but the individual values are characterized by a slightly increased variance [42]. The SW components of the upper half space are measured at roof height of the nearby building of the meteorological site.
At RWA, Tg from a black globe thermometer is also available, as are ta and va taken at the height of the black globe. tg and ta are quantified with copper-constantan thermocouples according to the standard TGL 32603/2 [43], where ta is additionally controlled against the air temperature measured with a standard PT-100 resistance thermometer [44]. va is taken with a Thies opto-electric anemometer, start-up-speed 0.3 (m/s). 10 min averages in UTC are accumulated to hourly means in true solar time.

2.6. Assessment Methodology and Reference

ISO 7726 [8] specifies the accuracy requirements for measuring methods based on the extent of the thermal annoyance being assessed. The more stringent requirements are for methods that relate to measurements carried out in moderate environments approaching comfort conditions, the less stringent are for environments subject to greater thermal stress. The latter may be considered for outdoor applications. Slightly simplified and thereby fractionally intensified in the extremes, the accuracy requirements of ISO 7726 [8] in Tmrt are ±2 K for comfort and ±5 K for thermal stress. Subsequently, methods for the calculation of Tmrt are tested against a reference method. It is assessed whether the differences in Tmrt against the reference is within the thresholds ±2 K or ±5 K. In order to comply with the ISO standard, the maximum error probability (α2K, α5K) has to be less than 5%. Further statistical values that are provided are the bias (K) as the mean of the differences, the root mean square error RMSE (K), and for linear regressions intercept (a) and slope (b), the coefficient of determination (R2) and the standard error (SE). SE is a measure of scattering of the sample points related to the regression value.
The assessment methodology requires a reference method. All three methods can be traced to proposals of ISO 7726 [8] for the measurement/calculation of Tmrt:
  • Tmrt,r,KMM: Calculation from the temperature of the surrounding surfaces (Equation (1)) in the form for outdoor use [1,12]. fp (Equation (2)) parameterized from diagrams [6]. This parameterization is confirmed by more recent measurements (Figure 1) [11]. In the extended form Tmrt,r,KMM accounts for fcs. Tmrt,r,KMM can be calculated based on standard meteorological radiant measurements.
  • Tmrt,r,6-Dir: Calculation from plane radiant temperatures [8] in the form for outdoor use [1,15]. The reference person is slightly simplified by a rectangular solid (Figure 1). The method implicitly accounts for fcs, because measured irradiances incident on vertical planes are used that depend on the azimuth of a plane relative to the azimuth of the solar disk. Variations in the accuracy of the instrument have been observed [45,46,47].
  • Tmrt,Tg,BG: Derived based on measured tg, ta and va (passive radiometry). ISO 7726 qualifies Tmrt,Tg,BG to be an approximation due to the difference in shape between a person and a globe. The vertical component in EI will be over-estimated in relation to that received by a standing person. The SW absorptivity significantly differs from that of the outer surface of a clothed person.
Tmrt,r,KMM is adopted as the reference because it applies directly observable radiant components (subscript: r) that are used as standard in meteorology. In Tmrt,r,6−Dir it is implicitly accounted for the circumsolar fraction of ED. However, this advantage cannot be completely used, because the data must be modeled (Section 2.5). Furthermore, fp for Tmrt,r,KMM is derived from precise measurements of the real human bodies [6,10,11], while in Tmrt,r,6-Dir the shape of the human body is slightly simplified to a rectangular solid.

3. Results

3.1. Six-Direction Measurements

Variations in the accuracy of 6-Dir measurements have been observed [45,46,47]. Because the 6-Dir method assumes the subject to be a rectangular solid, its vertical cross-section for unidirectional SW irradiance is a function of the difference in azimuth between the solar azimuth and the azimuth angle of one of the vertical surfaces (ΔAaz (degrees)) that are irradiated by the unidirectional fraction (1 − Df) in EG [30]. Df = (1 − fcs)∙ED/EG is the isotropic diffuse fraction. The unidirectional cross-section is smallest for ΔAaz = 0° because there is only one sunlit vertical surface, and greatest for 45°. In contrast, the vertical cross-section of a circular cylinder is azimuthally independent. Figure 1 presents the projected area factor of a cylinder that has the same relation of the cylindrical surface area and the surface area of the two parallel planes as the total vertical to the horizontal areas of the rectangular solid. Compared to that cylinder, the rectangular solid deviates in received irradiance due to the anisotropic fraction in SW irradiance (ΔEan (W/m2)) indicated by:
Δ E an = ( 1 D f ) · E G · cot ( γ s ) · ( ( 1 + sin ( 2 · Δ A az ) ) 1.273 )
Because they are based on the same measured radiant fluxes, differences in Tmrt,r,6-Dir and Tmrt,r,KMM can only arise from the anisotropic radiant parameters and the slightly differing fp (see Figure 1). Therefore, the evaluation of differences is restricted to daylight hours.
Figure 2a,b summarize the results for two azimuthal orientations of the 6-Dir instrument, where (a) assumes a measuring routine in which one of the vertical surfaces is set facing south, and (b) assumes that one of the vertical surfaces faces the solar azimuth. The reference accounts for variable fcs, despite the fact that it has to be modelled based on measurements. Figure 2c shows the effect of a reference that does not account for the circumsolar fraction in ED, i.e., in Equation (3) fcs = 0, as was originally suggested [12]. Figure 2d enables Figure 2a,b to be compared to a circular cylinder.
All orientations and assumptions remain within the ISO 7726 thresholds for thermal stress of ±5 (K) (Table 2). However, the comfort threshold of ±2 (K) is met only by the cylinder and with an error probability ≤1.3% by the 6-Dir instrument facing south. Because of its azimuthal dependence, the scattering in the tmrt differences is stronger for the rectangular solid than for the cylinder (Figure 2a,d). It significantly increases if the reference does not account for variable fcs (Figure 2c versus Figure 2a). If a measuring routine is assumed in which one of the vertical surfaces faces the solar azimuth (Figure 2b), this results in a significant negative bias of approx. −1 (K), because only one of the vertical surfaces is sunlit with unidirectional irradiance (Table 2). Thus, the bias due to the azimuthal error approaches zero if the instrument has a fixed orientation. The statistics of Table 2 include the data of the sites LER and SBO where the upward SW and LW components are modeled. The statistical results do not differ if the data of the two sites are not accounted for.
The third order polynomials in Figure 2 prove the impact of the more greatly simplified assumptions concerning the human body compared to fp of Equation (2). Especially for increased Tmrt,r,KMM, the differences tend toward values greater than zero. Here, the higher values are frequently combined with increased γs and, thus, with a higher probability of anisotropic SW irradiances. The upper surface of the rectangular solid as well as the cover of the cylinder represents head and shoulder girdle of a human less precisely than Fanger’s fp. As a consequence of the azimuthally dependent conditions of irradiance, the RMSE of the differences is almost double that in the rectangular solid compared to the cylinder (Table 2). In contrast, standard radiation measurements apply horizontal receivers that are by definition azimuthally independent. This gives additional support to the selection of Tmrt,r,KMM as the reference.
Figure 3 highlights the effect of circumsolar anisotropy in ED, fcs = variable, for determining Tmrt,r,KMM by comparing it with the case in which it is neglected, fcs = 0 [12]. In the extremes, Tmrt,r−fcs=var,KMM can deviate from Tmrt,r−fcs=0,KMM by approx. +5 (K) for solar altitudes γs close to the horizon and by approx. −7 (K) for the maximal γs of about 68 (degrees) within the data (due to SYNOPs only for 06 and 18 UTC at SBO (Table 1), higher solar elevations using modeled components are not attainable). Because the reference person is standing, the area projection factor fp for unidirectional SW irradiance increases with decreasing γs (Figure 1) and provides evidence for the positive deviations. For a given γs, the scattering in the differences arises from variable cloudiness due to the decreasing anisotropy effect with increasing cloud levels. At γs ≈ 32 (degrees), the optical air mass for the direct solar beam equals the mean value of the isotropic diffuse fraction and, thus, the anisotropic part in ED contributes to Tmrt,r,KMM by the same amount as if it were treated as isotropic. Because the optical air mass decreases in proportion to the increasing γs, there is a lower probability of the direct solar beam being scattered in its transfer to the earth’s surface. Hence, for high γs the differences tend towards a higher absolute value.

3.2. Black Globe Thermometer

A total of 25311 hourly mean values of measured Tg of a black globe is available for Warnemuende (RWA). They also include hours with precipitation that may impact Tg due to evaporative effects etc. Based on hourly synoptic observations, all Tg means are excluded that, in the present weather or within the past few hours, showed WMO-FM12 code numbers [48] greater than 10, i.e., cases with drizzle, rain, snow, thunderstorm etc. Furthermore, hours are excluded with an air speed at the height of the globe thermometer which is less than that of forced convection and/or with a difference in the air temperature registered by the thermocouples and the PT100 thermometer that exceeds 2 (K). Thus, 9191 values remain that are all covered by measured radiant components required to calculate Tmrt,r,KMM.
tg is not a thermodynamic property [49] but a measure of the complex radiant and convective effects of the environment on the black globe. Figure 4 examines some of them and their implications for the derived Tmrt,Tg,BG. It is also closely related to frequent attempts to predict tg based on standard meteorological measurements [34,49,50,51,52,53,54]. Figure 4a relates tg (°C) to ta (°C), where tg is split into three bins of radiant conditions: (i) at nighttime, i.e., sun is below the horizon, at daytime dependent on the value of EI: (ii) cloudy, i.e., EI < 120 (W/m2), no sunshine according to WMO [14], and (iii) sunny, i.e., EI ≥ 120 (W/m2).
At nighttime the radiant heat transfer is limited to EA and EE, the thermal spectral band. Then, ta and tg are closely and linearly regressed with the slope of the regression line close to unit (Table 3). The difference tgta is negatively biased by −0.9 (K), which provides evidence that, in most cases, it is the lower half space that in balance transfers radiant heat to the atmospheric environment. The scattering of the data can be attributed predominately to the varying va that is responsible for the convectional heat loss.
During the daytime, SW irradiation transfers heat onto the black globe, and the absorbed SW energy very often results in an increased tg, the amount of which depends on the SW intensity. At daytime in cloudy hours the sky is most frequently overcast, i.e., ED accounts for the largest share in EG. Compared to nighttime, the absorbed SW irradiance gives reason that the bias turns to positive and that the slope of the regression increases to 1.13. Compared to nighttime, the standard error is increased by a factor three and amounts to 1.9 (K). In the sunny hours, there is stronger SW heat transfer onto the black globe, the slope increases to 1.18, at +7.1 (K) the regression has a markedly positive bias and the standard error increases to +2.9 (K) (Table 3). The scattering in the data is significantly increased compared to the nighttime because va impacts on the convection, but tmrt also varies.
In the next step, a Tmrt,r,BG is simulated for a black globe based on measured radiant components. Because of the sphere’s simple geometry, its area projection factor is fp = 0.25 and independent of γs (Figure 1). Then, Tmrt,r,BG can be derived by applying Equation (3), assuming αsw,Sk = 0.95 and εsw,Sk = εg =0.95 [8]. Because it is based in this case on measured radiant quantities, Tmrt,r,BG is assumed as the reference. By highlighting the difference to Tmrt,r,BG, Figure 4b shows that Tmrt,Tg,BG significantly underestimates Tmrt,rKMM, bias −1.9 (K), which in this study is defined as reference (Table 4). The ISO 7726 [8] requirements for neither the comfort class nor thermal stress are met. The differences scatter significantly, especially for values of Tmrt,r,BG greater than approx. +20 (°C).
Figure 5a relates the difference of simulated Tmrt,r,BG to the reference Tmrt,r,KMM, where both Tmrt,r,BG and Tmrt,r,KMM use the same measured radiant components. The difference increases sharply with rising Tmrt,r,KMM because the SW absorption of the black globe is αsw,Sk = 0.95 compared to KMM of αsw,Sk = 0.70. There is noticeable data scatter around their 4th order polynomial, SE = 1.68 (K), due to the fact that fp of KMM is a function of γs while this is not the case for a sphere. Because of the enhanced SW absorption, the simulated Tmrt,r,BG does not meet the ISO requirements, either for the comfort or for the stress class (Table 4).
This alters when simulating a Tmrt,r,GG for a gray globe (GG), assuming αsw,Sk = 0.70 as for KMM (Figure 5b, Table 4). The fp-related scattering around the second-order polynomial remains unchanged; it is, however, somewhat reduced, SE = 1.25 (K). Because the simulated Tmrt,r,GG complies with the ISO requirements both for the comfort and for the stress class (Table 4), it becomes evident that the fp dependence is only of minor importance compared to the SW absorption. Time periods with summery more or less cloudless skies around noon may be characterized by EI > 600 (W/m2). Then, α5K is just beyond compliance with the ISO 7726 stress class due to the differences in shape between a globe and a standing person (Table 4).
Figure 5c compares Tmrt,Tg,BG to Tmrt,r,KMM: The scattering of the data values around their polynomial fits, SE = 4.66 (K), is considerably greater compared to Tmrt,r,BG (Figure 5a), SE = 1.68 K). In total, however, the values of the Tg-derived Tmrt,Tg,BG are lower than those of Tmrt,r,BG. This fact is clearly demonstrated by the two polynomial fits, where the dashed line is the fit of Figure 5a. Overall, Tmrt,Tg,BG does not comply with the ISO 7726 requirements, either for the comfort or for the thermal stress class (Table 4). Splitting the data into nighttime values and the daytime data into cloudy and sunny components reveals that the ISO 7726 class for thermal stress is met at nighttime (Table 4). There is a slight bias in the data, −0.27 (K), RMSE = 1.89 (K). During the daytime, the data do not comply with the ISO 7726 requirements, there is a positive bias for cloudy conditions +0.83 (K), increasing to +3.79 (K) during sunny periods (Table 4). Thus, the daytime results of the globe instrument show significant shortcomings compared to those derived from active radiometry. These shortcomings are attributable to the SW irradiation EG that covers the isotropic diffuse fraction of ED and the unidirectional fraction EI as well as the circumsolar part of ED. Especially under sunny conditions and high radiant intensity the unidirectional fraction accounts for the largest share in EG with ~75–85 %. Thus, EI is suitable to characterize conditions where the black globe thermometer is subject to strong asymmetric SW irradiation. Plotted over EI, the daytime data of the difference tmrt,Tg,BGtmrt,r,KMM show a bias of +2.2 (K) for 120 < EI ≤ 600 (W/m2) due to the enhanced SW absorption of the black globe compared to KMM (Figure 5d, Table 4). Stronger asymmetric radiant conditions can be attributed to values of EI > 600 (W/m2). Here the bias rises sharply with increasing EI (Table 4). For a gray globe that has the same radiant cross-section as the black globe but a reduced SW absorptance (αsw,Sk ~ 0.70), the maximally absorbed energy from unidirectional irradiances may be reduced to approx. 600 (W/m2) and the bias will be closer to zero. Because of the comparable amount of absorbed unidirectional radiant heat, it can be inferred from Figure 5d that the scattering in the data of a gray globe will be approximately the same as for the black globe in the region EI < 600 (W/m2), and, thus, significantly enhanced compared to Tmrt,r,GG (Figure 5b) and markedly too large to comply with the ISO 7726 stress class.

4. Discussion

At nighttime, Tmrt,Tg,BG remains within the ISO 7726 [8] thresholds for the stress class when applied as a measure for Tmrt,r,KMM. During daytime, and especially in summery sunny conditions, there is a significant positive bias and, in particular, markedly greater scattering in the values so that the stress class thresholds are far exceeded. This provides additional evidence that the increased scattering is a typical outdoor problem of a black globe thermometer, and coincides with other findings (e.g., [55]). Under direct sun, the SW irradiance is strongly asymmetric; a temperature difference of up to 7 (K) in opposing areas of the inner surface has been measured [56]. Additionally, the wind field on the outer border of the sphere’s film layer is not entirely uniform [57]. Therefore, the wind direction relative to the hot spot of direct irradiation can influence the overall convectional cooling of the globe. This spatial relationship between the fields of irradiance and of wind is important for the internal convection. Both theoretically and in experiments, the internal flow and temperature fields become asymmetric [37,55,58]. Thus, under strong asymmetric irradiation the temperature at the sphere’s center is unable to represent the average surface temperature of the globe thermometer system.
The black globe was originally introduced in order to measure radiant heat in relation to human comfort [17]. By chance, N u ¯ of the applied sphere with dg = 0.15 (m) is identical to that of a naked recumbent man [57]. Based on the similitude theory, this diameter is fundamental when Tg is related to human comfort, i.e., Tg is closely correlated with the operative temperature. Nevertheless, in the case of the black globe, it has the disadvantage that its SW absorption is significantly higher than that of human skin/clothing and thus tends to overestimate Tmrt when applied to human thermophysiology [22]. In contrast to comfort, the diameter is not predetermined using the black globe as a passive radiometer in order to determine Tmrt [16]. According to Equation (8), smaller diameters will improve the convective heat loss but will demand more accurate measurements of air velocity and ambient temperature [8,22,59,60].
Rewriting the transient form of the globe’s heat balance, Equation (6), in a linearized form allows, through temporal integration, the time constant of the instrument (τ (s)) to be determined for a stepwise change [38]: (QIrQc)outer = εg· h ¯ r , g ∙(TmrtTg) − h ¯ c , g ∙(TgTa) = −ρc∙δ∙dT/dt, where h ¯ r , g ≈ 4∙σSB∙((Tmrt + Tg)/2)3 (W∙m−2∙K−1) is the mean radiant heat transfer coefficient [61]. τ represents the time in which Tg reduces by 63.2% to the value of equilibrium. Within the response time, i.e., six to seven times τ, Tg has balanced out. Restricted to the heat exchange between the globe’s shell and the environment, τ is a function of h ¯ c , g . This depends, among other things, on the globe’s geometry and its material characteristics, as well as on h ¯ r , g . τ strongly decreases as the result of increasing h ¯ c , g that is a function of va. Outdoors, va is greater than in typical indoor conditions at va ≤ 0.2 (m/s). At RWA, where the measuring field is located in the dunes bordering the Baltic Sea, the mean of all records is va = 2.34 (m/s), the median 1.8 (m/s). In typical summery conditions, ta = 18.4 (°C), va = 2.6 (m/s), tmrt,Tg,BG = 25.8 (°C), the calculated response time is 5.25 (min) and is lower than for the same conditions by a factor of ~3, yet va = 0.20 (m/s). However, this does not take into account the heat conductivity of the shell’s material and especially not the radiant properties of the inner surface and the mass of the thermometer’s bulb which also affect the time constant of the instrument. The mass of the applied thermocouple is small, the emissivity of the oxidized copper inside the sphere is, at ~0.4, relatively low and there is only free convection. Hence, overall the instrument’s response time will be increased, but should not exceed observed values of about 10 min for outdoor air velocities of ~2 (m/s) [62]. Instruments have been constructed with a fan inside the sphere for forced convection in order to improve their response time [36,63]. However, they have not achieved greater acceptance.
Outdoors the environment is not in a steady state, i.e., the globe is impacted by short-term fluctuations, e.g., by wind gusts and brief cloud shadows in the case of broken cloud coverage, but usually this is only temporary. Hence, the hourly averages of measured Tg will rarely represent a steady state in the globe thermometer. Over the course of the day, they may include a time shift compared to the radiances, however these can be regarded as more or less insignificant.
Indoor black globe systems have been introduced, the shell of which consists of a table-tennis ball (acrylic), diameter ~0.04 (m), that is painted matt black. Their purpose is to achieve a quicker response time and lower instrument costs [64,65]. As a result of these reported advantages, such black globe instruments are now also being used outdoors [20,21]. However, under the strong asymmetric SW irradiation found outdoors, the heat conductivity of acrylic can no longer be ignored in terms of the heat balance of the system because it is lower than that of copper by a factor of ~1000. This may outweigh the reduced response time due to the lower mass density of acrylic compared to copper, and the reduced diameter. It is not recommended to use table-tennis balls outdoors as black globe thermometers in order to determine Tmrt because they yield unpredictable errors in the measurements [37,55].
The wet-bulb-globe-temperature (WBGT) index [18] is an index of thermal stress and has been adopted as standard ISO 7243 [19]. WBGT has the advantage that it can also be evaluated directly outdoors by taking readings of its three input components: the natural wet bulb temperature (tnw (°C)) as well as ta (°C) and tg (°C). They are summarized in the weighted average WBGT = 0.7∙tnw + 0.1∙ta + 0.2∙tg. The weights are empirically determined by linking them with heat casualties and work-rest cycles needed to prevent thermal strain on humans in the workplace [3,18,66,67]. Tmrt is not explicitly required as WBGT input. Hence, uncertainties in tg as found in this study are irrelevant for the accuracy of the WBGT index because they are accounted for in the weights. However, when the diameter of the globe used differs from the standard 0.15 (m), the measured tg value has to be adapted to the standard configuration [35,50,68]. Here the algorithm is based on va and the magnitude of the difference tgta [62]. It is difficult to measure tnw and tg, which are not thermodynamic properties [49]. Thus, measures have been undertaken in order to derive them from environmental data. In the case of tg, this can be achieved by solving Equation (7) for Tg. The required input Tmrt may be derived, assuming that it is a linear function of ta [51,52], i.e., radiant flux densities are only estimated. In more recent approaches, Tmrt is explicitly derived from readily available standard meteorological data that include the above-mentioned data from active radiometry (e.g., [24,34,53,54,67]). Here, Tmrt,r,BG can be calculated for a black globe by means of Equation (3) using fp = 0.25 and αsw,Sk = 0.95. Then, Equation (7) is solved for Tg by iteration or by using empirical/polynomial methods. As this study found, it is advantageous for such empirical approaches to differentiate during the daytime between conditions in which the black globe is sunlit or is shaded [50] and to address separately nighttime data that are more closely linked to ta (Figure 4a) [54,69]. From the correlation of modeled against observed tg [49,50,54] and from the difference Tmrt,Tg,BGTmrt,r,KMM as a function of tg [70], it is apparent that the observed tg exhibits the same problem as seen here, especially under sunny conditions (see Figure 4a,b, Table 4).
With the exception of EI, observed standard radiant components refer to horizontally mounted receivers. These measurements have the advantage that they do not depend on the solar azimuth. In contrast, the irradiance on the vertically mounted receivers of the 6−Dir instrument depends on the solar azimuth and the measured values consist of radiant components incident from the lower and the upper half space. If the irradiance incident from the ground can be regarded as Lambertian equivalent, which is generally the case, then the sum of EI and the circumsolar fraction of ED can be derived from the irradiance difference of the sunlit and the shaded receivers accounting for γs [47]. This implies, however, an environment free of significant horizon limitations [33]. The 6−Dir instrument is mainly intended for measurements in micro-structured urban environments. In, e.g., a narrow street canyon with the measuring site positioned closer to a building, the albedo of the building surfaces will significantly affect the measured diffuse SW irradiation, and the precondition for the method is violated.
In an urban environment, standard active radiometry combined with Equation (3) can also be used in order to derive Tmrt for a human bioclimatological assessment. Deploying a sun tracker to measure EI or ED may be costly. An alternative could be to deploy a rotating shadow band irradiometer. After correction for its deviation from the Lambertian response and for its nonuniform spectral response, the instrument delivers both EG and ED in moderate to good quality, which allows EI to be inferred [71,72]. Furthermore, the input values for Equation (3) may be measured at roof level, i.e., without horizon limitations [33], or taken from a nearby meteorological site. Then, the derived Tmrt,r,KMM is valid for the sky fraction seen from the point of assessment, the sky view factor, and the observed radiation flux densities can serve as a force in modeling the radiation transfer within the complex urban environment [21,73,74,75]. The observations of irradiation at meteorological sites frequently do not include all the components required to derive Tmrt,r,KMM. It has been shown that it is possible to calculate it with the required accuracy for the ISO 7726 thermal stress class if EG and a synoptic observation are available to serve as the base in order to model the missing radiant components [30,76].
As part of the model output, Numerical Weather Prediction (NWP) usually delivers in hourly resolution not only ta, ea (2 m above ground) and va (10 m above ground) but also gridded data of irradiation sums starting with the initialized analysis. Based on the example of the Deutscher Wetterdienst (DWD), EG, ED, and the LW irradiation balance EAEE may be available. The latter can be broken down using the forecast temperature of the ground surface or of the snow/ice surface, and converted using the Stefan–Boltzmann law to EE. The SW ground-reflected irradiation ER is derivable using the predicted ground albedo that varies over time, in particular due to snow cover or sea ice. Then, in NWP postprocessing, all input parameters are available to derive not only Tmrt,r,KMM via Equation (3) but also thermal indices as a basis for forecasting the spatial and temporal development of thermal health hazards (see e.g., [77]). Furthermore, test reference years derived from gridded high resolution datasets can also be helpful in the context of planning and health issues [78,79].

5. Conclusions

Among the required input parameters for rational thermal indices, Tmrt is the most complex. The quality of these indices and of the derived thermal assessment of the environment depends on the quality of the input. The requirements for Tmrt are specified by ISO 7726.
Based on active radiometry Tmrt accounts for the absorbed radiation from both the SW and LW spectral bands. Most of the radiant components are isotropic. Restricted to the SW band there are, however, also unidirectional irradiances: EI and under clear or partly obstructed skies a significant circumsolar fraction fcs in ED. fcs significantly impacts Tmrt with noticeably reduced values for high solar altitudes and increased for low elevations. Accounting for fcs is essential for the accuracy of Tmrt. An extension to an existing algorithm is provided in order to include fcs into the Tmrt calculation. We applied Fanger’s area projection factor fp, which is standard in determining the radiant cross-section of a person for unidirectional irradiance, because it is precisely aligned with measurements.
The method KMM processes radiant flux densities measured with receivers that are horizontally mounted. Such measurements are readily available from standard meteorologically radiant observations. EI and fcs can be traced to them. A parameterization of fp is applied that is confirmed by more recent measurements. Hence, KMM serves as reference in order to study the accuracy of the other methods.
The 6−Dir instrument posits the human as a rectangular solid. The radiant flux densities incident on the vertical planes are provided by vertically mounted receivers. The contribution of the unidirectional SW fraction on the vertical planes depends to a certain degree on the difference in azimuth angle between a plane and the solar disk. However, the produced scattering remains in the frame of the ISO 7726 comfort class. Additionally, Tmrt,r,6-Dir is in the second order, also varied due to the simpler geometry compared to a precisely aligned human subject. The measured fluxes on the vertical planes contain the impact of fcs, an advantage of the method. Measurements on vertical planes are rarely available, i.e., they are not standard and are elaborate and expansive.
Tmrt,Tg,BG is derived by solving the energy balance equation of the black globe thermometer. At nighttime, a Tmrt,Tg,BG can be derived that complies with the ISO 7726 stress class. During daytime, Tmrt,Tg,BG basically yields larger values than Tmrt,r,KMM due to its significantly increased SW absorption. An effect of secondary importance is that a globe is less appropriate for representing a standing body posture. Sunny summer conditions are linked to strong asymmetric SW irradiation. This results in strong scattering in the measured Tg and derived Tmrt,Tg,BG. Hence, outdoors in sunny conditions it is impossible to derive a Tmrt value of sufficient accuracy that could serve as input for the rational thermal indices. Also measurements with a gray globe under stronger asymmetric radiant conditions will not comply with the ISO 7726 stress class due to basic deficits of the globe thermometer.
Henceforth, Tmrt,r,KMM should serve as a reference in the scope of human biometeorology. It can be derived not only from standard, meteorologically observed radiant flux densities but also from the relevant NWP output and from partially modeled components that fill in for missing observed fluxes.

Author Contributions

Conceptualization, H.S. and A.M.; methodology, H.S.; software, H.S.; validation, H.S.; formal analysis, H.S.; investigation, H.S.; resources, H.S.; data curation, H.S.; writing—original draft preparation, H.S.; writing—review and editing, H.S. and A.M.; visualization, H.S.; supervision, A.M.; All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

The WMO Baseline Network (BSRN) at ETH Zürich granted Weihs, University of Natural Resources and Life Sciences, Vienna, access to the radiation data. This study constitutes one of a range of purposes for which the data were retrieved. We would like to thank the following for the radiation measurements: M. Maturilli (Alfred-Wegener-Institut, Bremerhaven) as well as J.B. Orbak (Norwegian Polar Institute, Longyearbyen) for the data of Ny Alesund, Norway, P. Fishwick (UK Met Office, Lerwick) for Lerwick/Shetland, United Kingdom, L. Vuilleumier (MeteoSwiss, Payerne) for Payerne, Switzerland, and A. Zangvil (Institute for Desert Research) for Sde Boqer, Israel.

Conflicts of Interest

The authors declare no conflict of interest.

List of Essential Symbols

SymbolUnitDescription
Df-isotropic diffuse fraction of ED
dgmdiameter of the globe’s shell
eahPawater vapor pressure
EAW/m2long-wave downward radiant flux density (atmospheric)
EEW/m2long-wave upward radiant flux density (earth surface)
EDW/m2short-wave downward diffuse radiant flux density
EGW/m2short-wave downward global radiant flux density
EIW/m2short-wave downward direct radiant flux density
ERW/m2short-wave upward surface reflected radiant flux density
fa-angle factor for isotropic irradiances, horizon unobstructed
fcs-anisotropic circumsolar fraction in ED
fp-projected area factor
h ¯ c , g W∙m−2∙K−1convective heat transfer coefficient averaged over the globe thermometer’s surface
h ¯ r , g W∙m−2∙K−1radiant heat transfer coefficient averaged over the globe thermometer’s surface
N u ¯ -mean Nusselt number of the globe thermometer
SrW/m2radiant energy absorbed by the 6−direction instrument
ta, Ta°C, Kambient temperature
tg, Tg°C, Kglobe temperature
tmrt, Tmrt°C, Kmean radiant temperature
Tmrt,r,6−DirKTmrt derived from radiant fluxes measured with the 6−direction instrument
Tmrt,r,BGKsimulated Tmrt of a black globe applying measured radiant fluxes and Equation (3) with fp = 0.25 and αsw,Sk = 0.95
Tmrt,r,GGKsimulated Tmrt of a gray globe applying measured radiant fluxes and Equation (3) with fp = 0.25 and αsw,Sk = 0.70
Tmrt,r,KMMKTmrt derived from radiant fluxes applying Equation (3)
Tmrt,Tg,BGKTmrt of a black globe derived applying Equation (8) using measured Tg, Ta, va
vam/sair velocity
Greek
α2K, α5K%error probability in the compliance with the accuracy in Tmrt according to the ISO 7726 comfort (±2 K) or stress (±5 K) class
αsw,Sk-short-wave absorption coefficient of skin/clothing
γsdegreesolar elevation
εlw,Sk-long-wave emissivity of skin/clothing
ρs-short-wave surface albedo
σSB5.67∙10−8 W∙m−2∙K−4Stefan–Boltzmann constant

References

  1. Guideline VDI 3787/Part 2. Environmental Meteorology, Methods for the Human Biometeorological Evaluation of Climate and Air Quality for Urban and Regional Planning at Regional Level: Part I: Climate. Kommission Reinhaltung der Luft im VDI und DIN-Normenausschuss KRdL. In VDI/DIN Handbuch Reinhaltung der Luft, Band 1B: Umweltmeteorologie; Beuth: Berlin, Germany, 2008.
  2. Staiger, H.; Laschewski, G.; Matzarakis, A. Selection of Appropriate Thermal Indices for Applications in Human Biometeorological Studies. Atmosphere 2019, 10, 18. [Google Scholar] [CrossRef] [Green Version]
  3. Parsons, K.C. Environments. In The Effects of Hot, Moderate, and Cold Environments on Human Health, Comfort and Performance; Taylor & Francis: New York, NY, USA, 2003; p. 518. ISBN 0-415-23793-9. [Google Scholar]
  4. Kantor, N.; Unger, J. The most problematic variable in the course of human-biometeorological comfort assessment—The mean radiant temperature. Cent. Eur. J. Geosci. 2011, 3, 90–100. [Google Scholar] [CrossRef] [Green Version]
  5. Bohnenkamp, H.; Pasquay, W. Untersuchungen zu den Grundlagen des Energie- und Stoffwechsels. III. Mitteilung: Ein neuer Weg zur Bestimmung der für die Wärme-strahlung maßgebenden Oberfläche des Menschen. Die ”mittlere Strahlungstemperatur“. Pflüger’s Archiv 1931, 228, 79–99. [Google Scholar] [CrossRef]
  6. Fanger, P.O. Thermal Comfort. In Analysis and Applications in Environmental Engineering; McGraw-Hill Company: Detroit, MI, USA, 1972; p. 244. [Google Scholar]
  7. Gagge, A.P.; Fobelets, A.P.; Berglund, P.E. A Standard Predictive Index of Human Response to the Thermal Environment. ASHRAE Trans. 1986, 92, 709–731. [Google Scholar]
  8. ISO. ISO 7726. Ergonomics of the Thermal Environment—Instruments for Measuring Physical Quantities, 2nd ed.; International Standard ISO 7726:1998 from Technical Committee ISO/TC 159; ISO: Berlin, Germany, 1998. [Google Scholar]
  9. ASHRAE. ASHRAE Handbook: Fundamentals, 8th ed.; American Society of Heating and Air-Conditioning Engineers: Atlanta, GA, USA, 2001; p. 29. ISBN 1-931862-70-2. [Google Scholar]
  10. Kubaha, K.; Fiala, D.; Toftum, J.; Taki, A.H. Human projected area factors for detailed direct and diffuse solar radiation analysis. Int. J. Biometeorol. 2004, 49, 113–129. [Google Scholar] [CrossRef]
  11. Park, S.; Tuller, S.E. Human body area factors for radiation exchange analysis: Standing and walking postures. Int. J. Biometeorol. 2011, 55, 695–709. [Google Scholar] [CrossRef]
  12. Jendritzky, G.; Menz, G.; Schmidt-Kessen, W.; Schirmer, H. Methode zur Raumbe-Zogenen Bewertung der Thermischen Komponente im Bioklima des Menschen (Fortgeschriebenes Klima-Michel-Modell); Akademie für Raumforschung und Landesplanung: Hannover, Germany, 1990; p. 80. ISBN 3888382076. [Google Scholar]
  13. Ohmura, A.; Dutton, E.G.; Forgan, B.; Fröhlich, C.; Gilgen, H.; Hegner, H.; Heimo, A.; König-Langlo, G.; McArthur, B.; Müller, G.; et al. Baseline Surface Radiation Network (BRSN/WRCP): New Precision Radiometry for Climate Research. Bull. Am. Meteor. Soc. 1998, 79, 2115–2136. [Google Scholar] [CrossRef] [Green Version]
  14. WMO. Guide to Meteorological Instruments and Methods of Observation, 2008th ed.; WMO No. 8; updated 2010; World Meteorological Organization: Geneva, Switzerland, 2010; ISBN 978-92-63-10008-5. [Google Scholar]
  15. Höppe, P. Ein neues Verfahren zur Bestimmung der mittleren Strahlungstemperatur im Freien. Wetter und Leben 1992, 44, 147–151. [Google Scholar]
  16. Bedford, T.; Warner, C.G. The globe thermometer in studies of heating and ventilation. Epidemiol. Infect. 1934, 34, 458–473. [Google Scholar] [CrossRef] [Green Version]
  17. Vernon, H.M. The measurement of radiant heat in relation to human comfort. J. Ind. Hyg. 1932, 14, 95–111. [Google Scholar]
  18. Yaglou, C.P.; Minard, C.D. Control of heat casualties at military training centers. AMA Arch. Ind. Health 1957, 16, 302–316. [Google Scholar] [PubMed]
  19. ISO. ISO 7243: Hot Environments—Estimation of the Heat Stress on Working Man, Based on the WBGT-Index (Wet Bulb Globe Temperature), 2nd ed.; Reference number ISO 7243; ISO: Geneva, Switzerland, 1989. [Google Scholar]
  20. Nikolopoulou, M.; Baker, N.; Steemers, K. Improvements to the Globe Thermometer for Outdoor Use. Archit. Sci. Rev. 1999, 42, 27–34. [Google Scholar] [CrossRef]
  21. Thorsson, S.; Lindberg, F.; Eliasson, I.; Holmer, B. Different methods for estimating the mean radiant temperature in an outdoor urban setting. Int. J. Climatol. 2007, 27, 1983–1993. [Google Scholar] [CrossRef]
  22. ASHRAE. ASHRAE Handbook, Fundamentals, 4th ed.; American Society of Heating, Refrigerating, and Air-Conditioning Engineers Inc.: Atlanta, GA, USA, 2001; p. 34. [Google Scholar]
  23. Perez, R.; Ineichen, P.; Seals, R.; Zelenka, A. Making full use of the clearness index for parameterizing hourly insolation conditions. Sol. Energy 1990, 45, 111–114. [Google Scholar] [CrossRef] [Green Version]
  24. Gaspar, A.R.; Quintela, D.A. Physical modelling of globe and natural wet bulb temperatures to predict WBGT heat stress index in outdoor environments. Int. J. Biometeorol. 2009, 53, 221–230. [Google Scholar] [CrossRef] [PubMed]
  25. Gueymard, C.A. Direct and indirect uncertainties in the prediction of tilted irradiance for solar engineering applications. Sol. Energy 2009, 83, 432–444. [Google Scholar] [CrossRef]
  26. Hodder, S.G.; Parsons, K.C. The effects of solar radiation on thermal comfort. Int. J. Biometeorol. 2007, 51, 233–250. [Google Scholar] [CrossRef]
  27. Batlles, F.J.; Olmo, F.J.; Alados-Arboledas, L. On shadowband correction methods for diffuse irradiance measurements. Sol. Energy 1995, 54, 105–114. [Google Scholar] [CrossRef]
  28. Sanchez, G.; Serrano, A.; Cancillo, M.L.; Garcia, J.A. Comparison of shadow-ring correction models for diffuse solar irradiance. J. Geophys. Res. 2012, 117, 8. [Google Scholar] [CrossRef] [Green Version]
  29. Ineichen, P.; Guisan, O.; Perez, R. Ground-reflected radiation and albedo. Sol. Energy 1990, 44, 207–2014. [Google Scholar] [CrossRef] [Green Version]
  30. Staiger, H. Die Strahlungskomponente im Thermischen Wirkungskomplex für Operationelle Anwendungen in der Human-Biometeorologie. Ph.D. Thesis, University of Freiburg, Freiburg, Germany, 2014; p. 287. [Google Scholar]
  31. Reindl, D.T.; Beckman, W.A.; Duffie, J.A. Diffuse fraction correlations. Sol. Energy 1990, 45, 1–7. [Google Scholar] [CrossRef]
  32. Muneer, T.; Gueymard, C.; Kambezidis, H. Solar Radiation and Daylight Models, 2nd ed.; Butterworth-Heinemann: Oxsford, UK, 2004; p. 345. ISBN 0 7506 5974 2. [Google Scholar]
  33. VDI 3786/Part 5: Environmental meteorology: Meteorological measurements: Radiation. Kommission Reinhaltung der Luft im VDI und DIN-Normenausschuss KRdL. In VDI/DIN Handbuch Reinhaltung der Luft, Band 1B: Umweltmeteorologie; Beuth: Berlin, Germany, 2015.
  34. Liljegren, J.C.; Carhart, R.A.; Lawday, P.; Tschopp, S.; Sharp, R. Modeling the wet bulb globe temperature using standard meteorological measurements. J. Occup. Environ. Hyg. 2008, 5, 645–655. [Google Scholar] [CrossRef] [PubMed]
  35. Graves, K.W. Globe thermometer evaluation. Am. Ind. Hyg. Ass. J. 1974, 35, 30–40. [Google Scholar] [CrossRef] [PubMed]
  36. Hellon, R.F.; Crockford, G.W. Improvements to the globe thermometer. J. Appl. Physiol. 1959, 14, 649–650. [Google Scholar] [CrossRef]
  37. Fontana, L. Experimental study on the globe thermometer behaviour in conditions of asymmetry of the radiant temperature. Appl. Therm. Eng. 2010, 30, 732–740. [Google Scholar] [CrossRef] [Green Version]
  38. Berglund, L.G. Radiation Measurement for Thermal Comfort Assessment in the Built Environment. In Thermal—Analysis-Human Comfort—Indoor Environments. In Proceedings of the Symposium Held at the National Bureau of Standards, Gaithersburg, MA, USA, 11 February 1977; National Bureau of Standards: Gaithersburg, MA, USA, 1977; pp. 117–130. [Google Scholar]
  39. Kuehn, L.A.; Stubbs, R.A.; Weaver, R.S. Theory of the globe thermometer. J. Appl. Physiol. 1970, 29, 750–757. [Google Scholar] [CrossRef]
  40. Stolwijk, J.A.J.; Hardy, J.D.D. Temperature regulation in man—A theoretical study. Pflügers Archiv 1966, 291, 129–162. [Google Scholar] [CrossRef]
  41. Gueymard, C.A.; Myers, D.R. Evaluation of conventional and high-performance routine solar radiation measurements for improved solar resource, climatological trends, and radiative modeling. Sol. Energy 2009, 83, 171–185. [Google Scholar] [CrossRef]
  42. Dehne, K.; Bergholter, U.; Kasten, F. IEA Comparison of Longwave Radiometers; Report No. IEA-SHCP-9F-3; Deutscher Wetterdienst: Hamburg, Germany, 1993; p. 70. [Google Scholar]
  43. Technische Normen, Gütevorschriften, Lieferungsbedingungen DDR (TGL) 32603/2: Arbeitshygiene, Mikroklima in Arbeitsräumen, Meßmethode; BBSR: Berlin, Germany, 1975.
  44. Töpfer, M.; Garbatzok, M.; Bendisch, F. Bioklimaprogramm—Datenerfassungsprogramm; Forschungsinstitut für Bioklimatologie, Meteorologischer Dienst der DDR: Berlin, Germany, 1983. [Google Scholar]
  45. Kantor, N.; Lin, T.-P.; Matzarakis, A. Daytime relapse of the mean radiant temperature based on the six-directional method under unobstructed solar radiation. Int. J. Biometeorol. 2014, 58, 1615–1625. [Google Scholar] [CrossRef]
  46. Chen, Y.-C.; Lin, T.-P.; Matzarakis, A. Comparison of mean radiant temperature from field experiment and modelling: A case study in Freiburg, Germany. Theor. Appl. Climatol. 2014, 118, 535–551. [Google Scholar] [CrossRef]
  47. Holmer, B.; Lindberg, F.; Rayner, D.; Thorsson, S. How to transform the standing man from a box to a cylinder—A modified methodology to calculate mean radiant temperature in field studies and models. In Proceedings of the 9th International Conference on Urban Climate (ICUC9), Toulouse, France, 20–24 July 2015; p. 6. [Google Scholar]
  48. WMO. Manual on Codes—International Codes—Volume I.1-Part A—Alphanumeric Codes; WMO-No. 306; World Meteorological Organization: Geneva, Switzerland, 2009; ISBN 978-92-63-10306-2. [Google Scholar]
  49. Buonanno, G.; Frattolillo, A.; Vanoli, L. Direct and indirect measurement of WBGT index in transversal flow. Measurement 2001, 29, 127–135. [Google Scholar] [CrossRef]
  50. Bernard, T.E.; Barrow, C.A. Empirical approach to outdoor WBGT from meteorological data and performance of two different instrument designs. Ind. Health 2013, 51, 79–85. [Google Scholar] [CrossRef] [Green Version]
  51. Sullivan, C.D.; Gorton, R.L. A method of calculation of WBGT from environmental factors. ASHRAE Trans. 1976, 82, 279–292. [Google Scholar]
  52. Azer, N.Z.; Hsu, S. OSHA heat stress standards and the WBGT index. ASHRAE Trans. 1977, 83, 30–40. [Google Scholar]
  53. Hunter, C.H.; Minyard, C.O. Estimating Wet Bulb Globe Temperature Using Standard Meteorological Measurements; WSRCMS-99-00757; Department of Energy (DOE), Office of Scientific and Technical Information (OSTI), Westinghouse Savannah River Company: Aiken, SC, USA, 1999. [Google Scholar]
  54. Turco, S.H.N.; da Silva, T.G.F.; de Oliveira, G.M.; Leitao, M.M.V.B.R.; de Moura, M.S.B.; Pinheiro, C.; da Silva Padilha, C.V. Estimating black globe temperature based on meteorological data. In Proceedings of the Livestock Environment VIII, Iguassa Falls, Brazil, 31 August–4 September 2008; p. 6. [Google Scholar] [CrossRef]
  55. Wang, S.; Li, Y. Suitability of acrylic and copper globe thermometers for diurnal outdoor settings. Build. Environ. 2015, 98, 279–294. [Google Scholar] [CrossRef]
  56. Bond, T.E.; Kelly, C.F. The globe thermometer in agricultural research. Agric. Eng. 1955, 36, 251–255. [Google Scholar]
  57. Büttner, K. Physikalische Bioklimatologie—Probleme und Methoden; Akademische Verlagsgesellschaft m.b.H.: Leipzig, Germany, 1938; p. 153. [Google Scholar]
  58. Kee, R.J.; Landram, C.S.; Miles, J.C. Natural convection of a heat-generating fluid within closed vertical cylinders and spheres. J. Heat Transfer. 1976, 98, 55–61. [Google Scholar] [CrossRef]
  59. Hey, E.N. Small globe thermometers. J. Phys. E Sci. Instrum. 1968, 1, 955–957. [Google Scholar] [CrossRef]
  60. Hatch, T.F. Design requirements and limitations of a single-reading heat stress meter. Am. Ind. Hyg. Assoc. J. 1973, 34, 66–72. [Google Scholar] [CrossRef]
  61. Fountain, M. Instrumentation for thermal comfort measurements: The globe thermometer; Internal study, Centre for the Built Environment, Series Indoor Quality (IEQ), University of California: Berkeley, CA, USA, 1987; Available online: http://escholarship.org/uc/item/1qx8c7sm (accessed on 24 July 2020).
  62. Juang, Y.-J.; Lin, Y.-C. The Effect of Thermal Factors on the Measurement of Wet Bulb Globe Temperature. J. Occup. Saf. Health 2007, 15, 191–203. [Google Scholar] [CrossRef]
  63. Vincent, D.F. Improvements in the globe thermometer. J. Hyg. Camb. 1939, 39, 238–243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Humphreys, M.A. The optimum diameter for a globe thermometer for use indoors. Ann. Occup. Hyg. 1977, 20, 135–140. [Google Scholar] [CrossRef] [PubMed]
  65. De Dear, R. Ping-pong globe thermometers for mean radiant temperatures. Heat. Vent. Eng. 1988, 60, 10–11. [Google Scholar]
  66. Epstein, Y.; Moran, D.S. Thermal comfort and the heat stress indices. Ind. Health 2006, 44, 388–398. [Google Scholar] [CrossRef] [Green Version]
  67. Lemke, B.; Kjellstrom, T. Calculating workplace WBGT from meteorological data: A tool for climate change assessment. Ind. Health 2012, 50, 267–268. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. D’Ambrosio Alfano, F.R.; Malchaire, J.; Palella, B.I.; Riccio, G. WBGT Index revisited after 60 years of use. Ann. Occup. Hyg. 2014, 58, 955–970. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Hardcastle, S.; Butler, K. A comparison of globe, wet and dry temperature and humidity measuring devices available for heat stress assessment. In Proceedings of the 12th US/North American Mine Ventilation Symposium, Reno, NV, USA, 9–11 June 2008; pp. 181–190, ISBN 978-0-615-20009-5. [Google Scholar]
  70. Kantor, N.; Kovacs, A.; Lin, T.-P. Looking for simple correction functions between the mean radiant temperature from the “standard black globe” and the “six-directional” techniques in Taiwan. Theor. Appl. Climatol. 2015, 121, 99–111. [Google Scholar] [CrossRef]
  71. Wilbert, S.; Kleindiek, S.; Nouri, B.; Geuder, N.; Habte, A.; Schwandt, M.; Vignola, F. Uncertainty of rotating shadowband irradiometers and Si-pyranometers including the spectral irradiance error. AIP Conf. Proc. 1734, 1734, 11. [Google Scholar] [CrossRef] [Green Version]
  72. Vuilleumier, L.; Felix, C.; Vignola, F.; Blanc, P.; Badosa, J.; Kazantzidis, A.; Calpini, B. Performance Evaluation of Radiation Sensors for the Solar Energy Sector. Meteorologische Zeitschrift 2017, 26, 485–505. [Google Scholar] [CrossRef]
  73. Lindberg, F.; Holmer, B.; Thorsson, S. SOLWEIG 1.0—Modelling spatial variations of 3D radiant fluxes and mean radiant temperature in complex urban settings. Int. J. Biometeorol. 2008, 52, 697–713. [Google Scholar] [CrossRef]
  74. Huttner, S.; Bruse, M. Numerical modeling of the urban climate—A preview on ENVI-MET 4.0. In Proceedings of the Seventh International Conference on Urban Climate (ICUC7), Yokohama, Japan, 29 June—3 July 2009; p. 4. [Google Scholar]
  75. Matzarakis, A.; Matuschek, O. Sky view factor as a parameter in applied climatology—Rapid estimation by the SkyHelios model. Meteorologische Zeitschrift 2011, 20, 39–45. [Google Scholar] [CrossRef]
  76. Weihs, P.; Staiger, H.; Tinz, B.; Batchvarova, E.; Rieder, H.; Vuilleumier, L.; Maturille, M.; Jendritzky, G. The uncertainty of UTCI due to uncertainties in the determination of radiation fluxes derived from measured and observed meteorological data. Int. J. Biometeor. 2012, 56, 537–555. [Google Scholar] [CrossRef] [PubMed]
  77. Pappenberger, F.; Jendritzky, G.; Staiger, H.; Dutra, E.; Di Giuseppe, F.; Richardson, D.S.; Cloke, H.L. Global forecasting of thermal health hazards: The skill of probabilistic predictions of the Universal Thermal Climate Index (UTCI). Int. J. Biometeorol. 2015, 59, 311–323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Krähenmann, S.; Walter, A.; Brienen, S.; Imbery, F.; Matzarakis, A. High-resolution grids of hourly meteorological variables for Germany. Theor. Appl. Climatol. 2018, 131, 899–926. [Google Scholar] [CrossRef] [Green Version]
  79. Rosenfelder, M.; Koppe, C.; Pfafferott, J.; Matzarakis, A. Effects of ventilation behaviour on indoor heat load based on test reference years. Int. J. Biometeorol. 2016, 60, 277–287. [Google Scholar] [CrossRef]
Figure 1. Projected area factor fp,: both Klima-Michel-Modell (KMM) [12] and Park and Tuller [11] for the reference subject and additionally for a cylinder and a globe.
Figure 1. Projected area factor fp,: both Klima-Michel-Modell (KMM) [12] and Park and Tuller [11] for the reference subject and additionally for a cylinder and a globe.
Atmosphere 11 00805 g001
Figure 2. Difference of 6-Dir instrument and cylinder to the reference tmrt,r,KMM. Green dotted lines delimit the ISO comfort class ±2 (K), red dashed lines the stress class ±5 (K). tmrt,r,KMM accounts for variable fcs: (a) the 6-Dir is set to face south; (b) the 6-Dir moves to face the solar disk; (c) 6-Dir faces south, however, Tmrt,r,KMM does not account for anisotropy in ED, i.e., fcs = 0; (d) for a cylinder with the same relation of barrel and covers, and tmrt,r,KMM accounts for variable fcs.
Figure 2. Difference of 6-Dir instrument and cylinder to the reference tmrt,r,KMM. Green dotted lines delimit the ISO comfort class ±2 (K), red dashed lines the stress class ±5 (K). tmrt,r,KMM accounts for variable fcs: (a) the 6-Dir is set to face south; (b) the 6-Dir moves to face the solar disk; (c) 6-Dir faces south, however, Tmrt,r,KMM does not account for anisotropy in ED, i.e., fcs = 0; (d) for a cylinder with the same relation of barrel and covers, and tmrt,r,KMM accounts for variable fcs.
Atmosphere 11 00805 g002
Figure 3. Solar altitude difference in Tmrt,r−fcs=var,KMM, that accounts for circumsolar anisotropy in diffuse irradiance, to the case Tmrt,r−fcs=0,KMM, that disregards the anisotropy effect.
Figure 3. Solar altitude difference in Tmrt,r−fcs=var,KMM, that accounts for circumsolar anisotropy in diffuse irradiance, to the case Tmrt,r−fcs=0,KMM, that disregards the anisotropy effect.
Atmosphere 11 00805 g003
Figure 4. (a) Globe temperature tg versus air temperature ta, (b) difference Tmrt,Tg,BG (K) derived from measured tg of a black globe instrument (BG) minus Tmrt,r,BG (K) calculated from measured shortwave (SW) and longwave (LW) irradiation for the assumption of a black globe versus tmrt,r,BG (°C), where green dotted lines delimit the ISO comfort class ±2 (K) and red dashed lines the stress class ±5 (K).
Figure 4. (a) Globe temperature tg versus air temperature ta, (b) difference Tmrt,Tg,BG (K) derived from measured tg of a black globe instrument (BG) minus Tmrt,r,BG (K) calculated from measured shortwave (SW) and longwave (LW) irradiation for the assumption of a black globe versus tmrt,r,BG (°C), where green dotted lines delimit the ISO comfort class ±2 (K) and red dashed lines the stress class ±5 (K).
Atmosphere 11 00805 g004
Figure 5. (a) Difference of simulated Tmrt,r,BG, αsw,Sk = 0.95, reference Tmrt,Tg,KMM and a polynomial fit of the data given by the gray dashed line; (b) same as (a), however, Tmrt,r,GG for a gray globe, αsw,Sk = 0.70; (c) difference of Tmrt,Tg,BG as measured by a black globe and reference Tmrt,r,KMM with polynomial fit, solid line, and the fit of (a), dashed gray line for comparison; (d) same as (c), however, restricted to daytime values and plotted over the direct irradiation EI (W/m2).
Figure 5. (a) Difference of simulated Tmrt,r,BG, αsw,Sk = 0.95, reference Tmrt,Tg,KMM and a polynomial fit of the data given by the gray dashed line; (b) same as (a), however, Tmrt,r,GG for a gray globe, αsw,Sk = 0.70; (c) difference of Tmrt,Tg,BG as measured by a black globe and reference Tmrt,r,KMM with polynomial fit, solid line, and the fit of (a), dashed gray line for comparison; (d) same as (c), however, restricted to daytime values and plotted over the direct irradiation EI (W/m2).
Atmosphere 11 00805 g005
Table 1. Available observed (■) and missing (—) irradiation data in hourly resolution; SYNOPs: daily start time (temporal resolution) end time; observed black globe (BG) temperatures including auxiliary quantities.
Table 1. Available observed (■) and missing (—) irradiation data in hourly resolution; SYNOPs: daily start time (temporal resolution) end time; observed black globe (BG) temperatures including auxiliary quantities.
ComponentNYALERRWAPAYSBO
EI
ED
EG
ER
EA
EE
SYNOP06(3)1800(1)2300(1)2300(3)2106+18
BG: Tg + ta, va
Table 2. Six-directional (6-Dir) instrument and cylinder: Statistics for the difference Tmrt,r,6-DirTmrt,r,KMM and error probability (α2K, α5K) for compliance with ISO 7726; n = 29,833 daylight hours.
Table 2. Six-directional (6-Dir) instrument and cylinder: Statistics for the difference Tmrt,r,6-DirTmrt,r,KMM and error probability (α2K, α5K) for compliance with ISO 7726; n = 29,833 daylight hours.
6-Dir InstrumentBias (K)RMSE (K)α2K (%)α5K (%)
fcs = varfacing south−0.081.391.3<0.1
fcs = varfacing hour angle−0.972.099.8<0.1
fcs = 0facing south−0.772.0018.30.1
fcs = varcylinder−0.050.74<0.1<0.1
Table 3. Black globe instrument (BG): linear regression of tg (C) versus ta (C).
Table 3. Black globe instrument (BG): linear regression of tg (C) versus ta (C).
Data FractionTest ValueReferenceBias (K)RMSE (K)a (°C)bSE (K)R2
nighttimetgta−0.901.09−0.931.010.620.99
daytime: cloudytgta+2.203.05+1.091.131.940.94
daytime: sunnytgta+7.117.83+4.811.182.910.92
Table 4. Statistics for the difference Tmrt from a globe minus Tmrt of a reference based on measured radiant components. First section: comparison of values derived from BG measurements versus simulated BG values; second section: simulated values for a black (BG) and a gray globe (GG) versus the KMM reference; third section: measured BG values versus the KMM reference for different data fractions. Error probabilities (α2K, α5K) for compliance with ISO 7726 [8].
Table 4. Statistics for the difference Tmrt from a globe minus Tmrt of a reference based on measured radiant components. First section: comparison of values derived from BG measurements versus simulated BG values; second section: simulated values for a black (BG) and a gray globe (GG) versus the KMM reference; third section: measured BG values versus the KMM reference for different data fractions. Error probabilities (α2K, α5K) for compliance with ISO 7726 [8].
Data FractionTest ValueReferenceBias (K)RMSE (K)α2K %α5K %
totalTmrt,Tg,BGTmrt,r,BG−1.904.7447.87.6
totalTmrt,r,BGTmrt,r,KMM+2.775.1847.07.1
totalTmrt,r,GGTmrt,r,KMM0.001.27<0.1<0.1
120 < EI ≤ 600 (W/m2)Tmrt,r,GGTmrt,r,KMM−0.612.0424.40.4
EI > 600 (W/m2)Tmrt,r,GGTmrt,r,KMM+0.613.1645.26.0
totalTmrt,Tg,BGTmrt,r,KMM+0.864.8549.58.8
nighttimeTmrt,Tg,BGTmrt,r,KMM−0.271.899.3<0.1
daytime: cloudyTmrt,Tg,BGTmrt,r,KMM+0.834.1451.610.5
daytime: sunnyTmrt,Tg,BGTmrt,r,KMM+3.798.8978.349.1
120 < EI ≤ 600 (W/m2)Tmrt,Tg,BGTmrt,r,KMM+2.198.4876.946.4
EI > 600 (W/m2)Tmrt,Tg,BGTmrt,r,KMM+5.879.4179.852.3

Share and Cite

MDPI and ACS Style

Staiger, H.; Matzarakis, A. Accuracy of Mean Radiant Temperature Derived from Active and Passive Radiometry. Atmosphere 2020, 11, 805. https://doi.org/10.3390/atmos11080805

AMA Style

Staiger H, Matzarakis A. Accuracy of Mean Radiant Temperature Derived from Active and Passive Radiometry. Atmosphere. 2020; 11(8):805. https://doi.org/10.3390/atmos11080805

Chicago/Turabian Style

Staiger, Henning, and Andreas Matzarakis. 2020. "Accuracy of Mean Radiant Temperature Derived from Active and Passive Radiometry" Atmosphere 11, no. 8: 805. https://doi.org/10.3390/atmos11080805

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop