<<

Rabi et al.

Human Thermal : an José A. Rabi Irreversibility-Based Approach [email protected] University of São Paulo Emulating Empirical Clothed-Body Faculty of Animal Science and Food Engineering 13635-900 Pirassununga, SP, Correlations and the Conceptual

Robson L. Silva Energy Balance Equation [email protected] Federal University of Dourados Exergetic analysis can provide useful information as it enables the identification of irreversible phenomena bringing about entropy generation and, therefore, exergy losses Faculty of Engineering / Energy Engineering (also referred to as irreversibilities). As far as human thermal comfort is concerned, 79804-970 Dourados, MS, Brazil irreversibilities can be evaluated based on parameters related to both the occupant and his surroundings. As an attempt to suggest more insights for the exergetic analysis of thermal Celso E. L. Oliveira comfort, this paper calculates irreversibility rates for a sitting person wearing fairly light clothes and subjected to combinations of ambient air and mean radiant temperatures. The [email protected] thermodynamic model framework relies on the so-called conceptual energy balance University of São Paulo equation together with empirical correlations for invoked thermoregulatory Faculty of Animal Science and Food Engineering rates adapted for a clothed body. Results suggested that a minimum irreversibility rate 13635-900 Pirassununga, SP, Brazil may exist for particular combinations of the aforesaid surrounding temperatures. By separately considering the contribution of each thermoregulatory mechanism, the total irreversibility rate rendered itself more responsive to either convective or radiative clothing-influenced heat transfers, with exergy losses becoming lower if the body is able to transfer more heat (to the ambient) via . Keywords: , exergetic analysis, thermal comfort,

1 Introduction where actual and comfort entropy generations ( Sgen,act and Sgen,com ) are calculated based on occupants’ responses and on ambient Thermal comfort analyses may contribute to design bioclimatic variables. Referred to as human coefficient, the dimensionless and energy-efficient buildings as well as to outline safety (or health) parameter H accounts for variations due to factors like age, sex and precautions or standards for workers at food processing areas or race. Bearing in mind ASHRAE (or ISO 7730) definition for chambers. Several parameters have been defined to thermal comfort, namely, “that condition of mind which expresses assess human thermal sensation, acceptability or comfort (Givoni, satisfaction with the thermal environment” (ASHRAE, 2001), one 1981; Parsons, 1993) and they are referred to as ( i) direct, if based may regard the coefficient H as an attempt to link physiological to on data read from instruments used to imitate body responses, ( ii ) psychological responses, which may vary from person to person. empirical, when obtained via numerical regression of human Such an idea of introducing particular corrections was similarly physiological responses as occupants become exposed to distinct conceived by Humphreys and Hancock (2007) via an ‘adjusted ambient conditions, or ( iii ) rational, when based on theoretical thermal sensation’ concept to indicate how much occupant’s actual reasoning. thermal sensations exceed the desired ones on ASHRAE scale for Rational parameters have been usually defined in terms of an thermal comfort. energy balance applied to the human body, accounting for First and second laws of thermodynamics can be suitably concurrent effects from the well-known six basic factors, namely: combined to improve process efficiency. Resulting from such air temperature, radiant temperature, air , air speed, combination, exergy is a property that can be interpreted as the activity-related , and clothing (Fanger, 1970; Parsons, useful energy available for a system (Kotas, 1985; Bejan, 1988; 1993). In other words, rational parameters have evoked the first law Szargut, Morris and Steward, 1988). Taking into account parameters of thermodynamics to deal with thermoregulatory responses related related to the system as well as to its surroundings, exergetic to energy interactions between the human body and ambient analysis may identify dissipative or irreversible phenomena and parameters (the first four basic factors in the aforesaid list) compare a given process to the most efficient way by which it could combined with behavioural parameters (the last two factors in the be carried out, namely, an ideal reversible process. list). Yet, the first law cannot identify imperfections in real Links between thermoregulation and exergy have already been processes while it makes no distinction between heat and work attempted. Considering a two-compartment (core and skin) model interactions. for the human body, Prek (2004, 2005, 2006) and co-worker (Prek Depending on the amount of energy that can be transformed into and Butala, 2010) discussed exergy consumption rates with respect useful work, the second law introduces a qualitative feature to to related values for standard predicted mean vote (PMV), based on energy interactions. Taking into account the second law, the so- one-compartment model (Fanger, 1970), and for its modified value called objective thermal comfort index (OTCI) has been put forward (PMV *), proposed for any dry or humid environment (Gagge, as “the percentage deviation in the value of entropy generation from Stolwijk and Nishi, 1971). the comfort or equilibrium condition” (Boregowda, Tiwari and By the same token, one might attempt to correlate occupant’s Chaturvedi, 2001): thermal discomfort and exergy losses, equally referred to as irreversibilities. Based on a steady-state model framework and = − × OTCI (%) H (1[ Sgen, act Sgen, com )] 100 (1) preliminary concepts as discussed in a previous work (Rabi, Fresia and Benzal, 2006), the present work puts forward more insights to the exergetic analysis of human thermal comfort by relying on the so-called conceptual energy balance equation together with existing Paper received 27 November 2009. Paper accepted 30 June 2012 empirical correlations for the invoked thermoregulatory heat transfer Technical Editor: José Parise

450 / Vol. XXXIV, No. 4, October-December 2012 ABCM Human Thermal Comfort: an Irreversibility-Based Approach Emulating Empirical Clothed-Body Correlations… rates expressed in terms of a clothed body. Bearing in mind a com = comfort value (level) for entropy generation prospective rational assessment of thermal comfort, irreversibilities cond = conductive heat transfer are calculated and discussed for a sitting person wearing relatively conv = convective heat transfer light clothes and subjected to some combinations of ambient air and evap = evaporative heat loss mean radiant temperatures. gen = generation (due to imbalance) Another approach to human thermal comfort is given by Batato in = inlet (or input) of open system et al. (1990), where three different temperature levels are modelled isol = isolated system and submitted to an exergetic analysis. The human body was met = metabolic value considered at rest, which corresponds to the same condition moist = moisture (in air exhaled from lungs) considered herein, i.e., a sitting person wearing light clothes. musc = muscular value Data from Simone et al. (2011) point out a second-order out = outlet (or output) of open system polynomial relationship between human-body exergy consumption rad = radiative heat transfer rates, i.e., irreversibility rates, with a slightly cool sensation, resp = respiration value reaching minimum values when thermal sensation is close to sat = saturation value (water vapour) neutrality. The study considered sedentary people, male and female, skin = skin value subjected to convective and radiant heat transfers. sw = sweat (evaporation from the skin) In a recent work (Mady et al., 2012), the thermal behaviour of vap = water vapour (diffusion through the skin) human body was simulated by dividing it into 15 compartments. 0 = surrounding (ambient) reference value Exergy balances were accomplished considering radiation, Acronyms convection, evaporation and respiration and results indicated that human body becomes more efficient and destroys less exergy for EMR = exit matter reservoir lower relative humidity and higher temperatures. Mady et al. (2012) MER = mechanical energy reservoir elaborated a human thermal model for each part of the human body TER = thermal energy reservoir and considered thermoregulation and passive systems for a healthy person under basal conditions. Their results indicate that minimum Theory entropy production is achieved for an adult person and that exergy efficiency decreases during lifespan. Preliminary concepts and ideas Homeothermy is markedly characterized by energy interactions Nomenclature between body and ambient (Ivanov, 2006). Human body o 2 temperature should remain within a narrow range: between 36.7 C A = area, m o and 37.0 C, as measured in the mouth, while rectal temperature is E = energy content, J o about 0.6 C higher (Guyton, 1995). In line with the heat-dependent f = clothing surface area factor, dimensionless thermoregulation rationale (Webb, 1995) and pointing to the limited H = human coefficient, dimensionless efficiency of thermoregulatory mechanisms, Ivanov (2006) h = heat transfer coefficient (for convection or thermal discussed the responsiveness of the human thermoregulatory system 2⋅ radiation), W/(m K) with respect to body’s energy content and mean temperature. I = irreversibility (= exergy loss), J Assuming that thermoregulation attempts to follow an energy- = irreversibility rate, J/s = W efficient path, one may, in principle, argue whether the body is " 2 = irreversibility rate per unit of area, J/(s ⋅m) = W/m equally sensitive to energy quality and enquire about its 2 i = insulation (= thermal resistance), m ⋅K/W (or Clo = “preference” for a given energy interaction rather than another in 0.155 m 2⋅K/W) order to achieve homeothermy. = mass flow rate, kg/s An exergetic analysis of thermal comfort may indicate qualitative " = mass flow rate per unit of area (= mass flux), kg/(s ⋅m2) differences among distinct thermoregulatory heat transfers and one P = partial pressure (water vapour in air at saturation may start such analysis by discussing the following: condition), kPa • System definition : Rigorously, the body is a control volume Q& = heat transfer rate, J/s = W (open system) as it transfers water to the ambient as sweat and/or vapour through the skin (evapotranspiration) and as " = heat transfer rate per unit of area (= heat flux), J/(s ⋅m2) = moisture in exhaled air (respiration). Food and water intakes W/m 2 play an opposite role. Yet, depending on the situation (e.g., short ⋅ = entropy (generation) rate, J/(s K) = W/K time exposure), the amount of mass transferred can be as small s = specific entropy, J/(K ⋅kg) o as to be practically disregarded so that the body can be treated as T = temperature, K (or C) a control mass (closed system). t = time, s • Process dynamics and irreversibility : One may argue to what v = velocity (air speed), m/s extent thermoregulation entails quasi-equilibrium or non- = mechanical (muscular) work rate, J/s = W equilibrium processes, besides discussing whether or not they " = mechanical (muscular) work rate per unit of area, are steady-state processes. Comprising dissipative phenomena 2 2 J/(s ⋅m ) = W/m (direct dissipation of work into internal energy) or spontaneous Greek Symbols non-equilibrium processes (natural tendency of systems to achieve equilibrium with their surroundings), irreversibilities φ = air relative humidity, dimensionless always occur in real processes. Depending on the activity, Subscripts muscular (mechanical) work varies from about zero up to 25% act = actual value (level) for entropy generation of the total metabolic energy release (Parsons, 1993), with the air = surrounding (ambient) air resulting excess being transferred to the ambient as heat, over a body = human body (or body core) finite temperature difference. cl = clothing-influenced (clothing-related) value

J. of the Braz. Soc. of Mech. Sci. & Eng. Copyright  2012 by ABCM October-December 2012, Vol. XXXIV, No. 4 / 451 Rabi et al.

• Surroundings : With respect to the body, the ambient may behave ( ′′ − ′′ )+ ( ′′ + ′′ + ′′ )+ as a thermal energy reservoir (TER) for heat transfers and as an q&met w& musc q&evap, skin q&conv, skin q&rad, skin exit matter reservoir (EMR) for water transfers. Regarding (4) + ()′′ + ′′ = muscular work (if any), the concept of mechanical energy q&evap, resp q&conv, resp 0 reservoir (MER) can be considered. Irrespective of its interactions with the occupant, changes in the thermodynamic state of the ′′ Respiration heat transfer via thermal radiation is null ( q&rad, resp = 0) ambient are only due to variations in the external conditions. as humid air is assumed transparent. Fanger (1970) proposed to modify Eq. (4) to some extent, based Human thermoregulation: conceptual heat balance equation on the subsequent premises: and empirical correlations • Heat balance prevails so that body temperature is steady (homeothermy); and The first law of thermodynamics expresses an energy balance of • Sweat rate and mean are within comfort limits. a system over its work and/or heat interactions with the ASHRAE (or ISO 7730) has equally followed such thermal surroundings regardless of how much heat can be converted into comfort approach (ASHRAE, 2001). Further modifications in Eq. useful energy (work is useful energy by definition). Based on a one- (4) are such that ( i) skin-related heat transfers q′′ are split up compartment (one-node) approach for the human body, one may &evap, skin ′′ ′′ write the following energy balance (Bligh, 1985): into q&vap, skin , due to water vapour diffusion, and q&sw, skin , due to

sweat evaporation, while ( ii ) q&′′ and q&′′ become dE conv, skin rad, skin body = ()()− + + + + ′′ ′′ Q&met W&musc Q&evap Q&cond Q&conv Q&rad (2) q& and q& , respectively, i.e., convective and radiative heat dt conv, cl clrad, transfers account for the clothing influence. In view of that, the In line with the so-called heat-engine sign convention, in the conceptual energy balance equation becomes: > previous balance equation any Q& 0 indicates a heat gain by the ( ′′ − ′′ )+ ( ′′ + ′′ + ′′ + ′′ )+ q&met w& musc q&vap, skin q&sw, skin q&conv, cl q& clrad, body while any Q& < 0 refers to a heat loss. One should be aware (5) + ()′′ + ′′ = that a different sign convention is proposed for Eq. (2) in q&evap, resp q&conv, resp 0 (ASHRAE, 2001). In Eq. (2), energy accumulation is lumped into a sole system Following ASHRAE sign convention, Parsons (1993) presented (one-node) so that d Ebody /dt > 0 leads to a body temperature rise empirical correlations for the heat transfers considered in Eq. while d Ebody /dt < 0 results in a temperature drop. Referred to as net (5). Yet, as this work follows the heat-engine sign convention, a − minus sign must be properly introduced so that a negative value for heat release rate, the difference (Q&met W&musc ) gives the available q&′′ refers to a heat transfer from the occupant to the ambient. The energy from the total metabolic release ( Q&met > 0), after muscular aforementioned empirical correlations then become: ≤ power ( W&musc > 0) has been discounted. As W&musc 25.0 Q&met − ′′ = − ′′ − ′′ − (6a) (Parsons, 1993), such difference is always positive. The summation q&vap, skin 733.5[05.3 .0 00699 (q&met w& musc ) P airsat, ] + + + − ′′ = ′′ − ′′ − (6b) ( Q&evap Q&cond Q&conv Q&rad ) comprises evaporative, conductive, q&sw, skin [(42.0 q&met w& musc ]15.58) − ′′ = − convective, and radiative heat transfer rates, respectively. According q&conv, cl cl hf conv (Tcl Tair ) (6c) to the heat-engine sign convention applied to Eq. (2), negative − ′′ = × −8 + 4 − + 4 values for those rates denote heat transfers from the body to the q& clrad, 96.3 10 fcl [( Tcl )15.273 (Trad ])15.273 (6d) ambient. − ′′ = ′′ − q&evap, resp .0 00173 q&met 867.5( P airsat, ) (6 e) One may normalize Eq. (2) over different body sizes via its − q ′′ = 0014.0 q ′′ 34( −T ) division by the body surface area, Abody . If steady-state (homeothermy) &conv, resp &met air (6f) is assumed, one may then impose d( Ebody /Abody )/d t = 0 in order to obtain the so-called conceptual heat balance equation (Parsons, 1993) Similar expressions for the respiration-related heat fluxes are also per unit of body area: found in Prek (2006). 2 In the equations (6), all heat fluxes q&′′ are in W/m provided (q′′ − w′′ )+ (q′′ + q′′ + q′′ + q′′ ) = 0 (3) &met & musc &evap &cond &conv &rad that all temperatures (namely, air temperature Tair , mean radiant temperature Trad and clothed-body surface temperature Tcl ) are in °C ′′ = ′′ = while water vapour pressure Psat,air in ambient air (at the saturation where w& musc W&musc / Abody and all q& Q& / Abody terms have condition) is in kPa. Referred to as clothing area factor, fcl is a × −1 × −1 dimensions of energy time area . Though muscular power dimensionless parameter that depends on the icl ′′ w& musc could, in principle, comprise involuntary motions (e.g., and it can be assessed as (Parsons, 1993): peristalsis or heart beating) as well as voluntary motions (e.g., writing = + 2⋅ ⋅ −1 or walking), the former have already been taken into account in fcl 31.01 icl (icl in Clo units, 1 Clo = 0.155 m K W ) (7) ′′ experiments measuring the metabolic heat release q&met (Nishi, 1981). In the second pair of brackets of Eq. (3), conductive heat transfer A different correlation for fcl is found in Olesen (1982). The ′′ q&cond is usually neglected (Parsons, 1993) while evaporative, convective heat transfer coefficient hconv can be assessed from Tcl convective, and radiative terms are collected into heat transfers and Tair together with the air speed vair as (Olesen, 1982): ′′ through two separate pathways: ( i) skin ( q&skin ) and ( ii ) lungs due to = ( − 25.0 5.0 ) ⋅ −2⋅ −1 ′′ hconv max (38.2 Tcl Tair ) 1.12, vair (in W m K ) (8) respiration ( q&resp ). Accordingly, one may rewrite Eq. (3) as the following steady-state balance equation for the occupant’s body:

452 / Vol. XXXIV, No. 4, October-December 2012 ABCM Human Thermal Comfort: an Irreversibility-Based Approach Emulating Empirical Clothed-Body Correlations…

A simpler correlation based solely on vair is presented in Parsons (1993) for a seated person. In order to obtain the clothed-body surface temperature Tcl an iterative calculation procedure is here followed. Firstly, the sum ′′ + ′′ q&conv, cl q& clrad, is assessed by inserting in Eq. (5) all heat fluxes ′′ q& that are independent from Tcl in Eqs. (6). From a trial value ′′ assigned to Tcl , the radiative heat flux q& clrad, is assessed from Eq. (a) ′′ (6 d), which thus allows one to obtain q&conv, cl . With the help of Eqs.

(6 c) and (8), a new Tcl value is determined and the procedure is repeated until convergence is achieved. In the light of Eq. (5), such converged value for Tcl is exactly the one that satisfies the heat balance equation, thus rendering steady-state (i.e., homeothermy) a reasonable assumption. Conversely, a temperature Tcl different from such value leads to an energy imbalance, d( Ebody /Abody )/d t ≠ 0, so that the body should undergo either a temperature rise or drop.

Human thermoregulation: irreversibility (exergy loss) analysis Like entropy, exergy is exempt from a conservation law in the sense that an exergy-loss term must be introduced to close exergy balances. Also referred to as irreversibility, such term refers to the (b) degraded useful energy when real processes are carried out. By taking into account the conceptual energy balance equation, Eq. (5), and auxiliary empirical correlations, Eqs. (6) to (8), irreversibilities can be assessed under two distinct approaches with respect to occupant’s interactions via skin and lungs (respiration) with the ambient. Under the first approach, the occupant behaves as a control volume (open system) so that moisture and/or vapour losses are treated as mass flows, indicated as double-line arrows in Fig. 1(a). Under the second approach, the occupant behaves as a Figure 1. Sketch of the two possible model frameworks for irreversibility calculations: (a) occupant as a control volume (open system) or (b) control mass (closed system) and the effects resulting from those occupant as a control mass (closed system). mass transfers are modelled as heat transfers, Eqs. (6a), (6b) and (6e), indicated as single-line arrows in Fig. 1(b). Regardless of the approach followed, it is herein assumed that: Normalized to the body surface area Abody , one may assess the • Ambient and occupant form an isolated system and the latter is irreversibility rate (= exergy loss per unit of time) per unit of area undergoing a steady-state process; under the first approach as (Kotas, 1985):

• Heat-engine sign convention is adopted;  ′′  • No EMR is attributed for water / food intakes as occupant is ′′ = ′′ − ′′ − q& I& T0  ∑ m& out sout ∑ & sm inin ∑  (9) presumably not eating or drinking;  T  • ′′ EMR EMR TER Muscular power per unit of body surface area w& musc (if any) ′′ only comprises voluntary motions (as previously discussed) and where T0 is ambient temperature, m& and s are respectively the it is always from occupant to ambient, the latter behaving as a mass flux and the specific entropy of every matter stream crossing MER (the direction of such energy interaction remains fixed as the system boundary, and each q&′′ is a heat flux interaction with a indicated by the arrowheads in Fig. 1); • ′′ given TER at temperature T. One may follow a procedure similar to Direction of metabolic heat release remains fixed ( q&met > 0), as Prek (2004, 2005, 2006) so as to estimate outlet mass fluxes indicated by arrowheads in Fig. 1, so that the occupant always ′′ ′′ ′′ m& vap, skin , m& sw, skin and m& evap, resp by using the corresponding heat receives energy from a TER presumably at temperature T ; body ′′ ′′ ′′ • Directions of evaporative mass transfers (first approach) or fluxes q&vap, skin , q&sw, skin and q&evap, resp , given by Eqs. (6a), (6b) related heat fluxes (second approach) remain fixed from and (6e), together with the specific for liquid and vapour occupant to ambient, as indicated by the arrowheads in Fig. 1 phases (obtained via steam tables) at the thermodynamic conditions (under the first approach the ambient behaves as an EMR while of skin and lungs. Specific entropies ssw,skin , svap,skin and sevap,resp at it becomes a TER under the second); the outlet can be similarly obtained and it is assumed that • In contrast, convective heat transfer via respiration q′′ as ′′ = &conv, resp ∑ & sm inin 0 since water or food intakes are both disregarded. ′′ ′′ EMR well as clothing-related heat fluxes q&conv, cl and q& clrad, can be Yet, the present work follows the second approach, which either from occupant to ambient or in the opposite direction; suppresses all aforesaid exit mass fluxes by assuming that the specifically, a common TER is assigned to the convective occupant behaves as a control mass instantaneously or for a short ′′ ′′ transfers q&conv, resp and q&conv, cl (namely, the ambient air at Tair ) period of time. Each omitted mass flux is replaced (in while a distinct TER (at Trad ) is appointed to the radiative compensation) by a heat transfer with a TER assigned at ′′ transfer q& clrad, . temperature Tair , as the suppressed EMR referred to ambient air. Accordingly, Eq. (9), under the second approach, renders itself to:

J. of the Braz. Soc. of Mech. Sci. & Eng. Copyright  2012 by ABCM October-December 2012, Vol. XXXIV, No. 4 / 453 Rabi et al.

 q ′′ q ′′ q ′′ One could be tempted to wonder about a probable scenario ′′ = − &vap, skin − &sw, skin − &evap, resp − I& T0 where Tcl > Tbody with the body (or only part of it) behaving as a heat  Tair Tair Tair pump. Accordingly, one should then infer about the temperature (10) differences as well as the source for and the nature of the necessary q ′′ q ′′ q ′′ ′′  − &conv, resp − &conv, cl − & clrad, − q&met  work input. Furthermore, one may also abandon the steady-state  Tair Tair Trad Tbody  assumption for thermal comfort (Fanger, 1970) and put forward a transient model. By acknowledging that thermal sensations (e.g., thermal pleasure) are transient in nature and cannot be experienced Results and Discussion under steady-state conditions, Parsons (1993) regarded thermal A sitting person (occupant) wearing rather light clothes was comfort simply as “a lack of discomfort”. One could also propose a preliminary considered and Table 1 presents the values assigned to two-node (e.g., body-clothing) model framework, but any of those previous issues are beyond the scope of this work. some necessary thermal comfort parameters. By inserting those By interpreting T as the clothing temperature, one may suggest values in Eqs. (6 b) and (7), one obtains some initial results (also cl ′′ < ′′ < that any heat loss q&conv, cl 0 or q& clrad, 0 first occurs from body presented in Table 1), which are independent from temperatures Tair and Trad . to clothing over a temperature difference Tbody − Tcl > 0 and then transferred from clothing to ambient over a difference Tcl − Tair > 0 or Tcl − Trad > 0, depending on the heat transfer mechanism. In order Table 1. Thermal comfort parameters and initial results independent of to prevent violations of the second law of thermodynamics (i.e., heat temperatures Tair and Trad . being spontaneously transferred from a colder system to a hotter one), care was exercised to only consider allowable combinations of Thermal comfort parameter or Value assigned / calculated Tair and Trad . For some mean radiant temperature values (namely, initial result o o o o o o Trad = 16 C, 18 C, 20 C, 22 C, 24 C and 26 C) and using a proper ≤ ≤ ′′ 2 variation range for the air temperature (namely, 17°C Tair 33°C), Metabolic heat release flux per unit of q&met = 100 W/m body area Fig. 2 shows the converged values for the clothed-body surface temperature Tcl , which seems to vary almost linearly with ′′ temperature T . Muscular power per unit of body area w& musc = 0 air

2 Clothing insulation icl = 0.6 Clo = 0.093 m ⋅K/W 38 36 Ambient air velocity vair = 0.25 m/s 34 Ambient air relative humidity φ = 1 (100%) 32 o Body (core) temperature Tbody = 37.0 C

(°C) 30

cl T_rad = 26°C ′′ − 2 T Heat flux due to sweating (from q&sw, skin = 17.577 W/m 28 T_rad = 24°C occupant to ambient) T_rad = 22°C 26 T_rad = 20°C Dimensionless clothing area factor f = 1.186 cl 24 T_rad = 18°C T_rad = 16°C According to model equations presented in section 2, 22 17 19 21 23 25 27 29 31 33 subsequent results depend on either T or T (or both) and care air rad T (°C) should be exercised to fulfil the applicability of the aforementioned air ′′ Figure 2. Clothed-body surface temperature Tcl (converged value) as a correlations. One constraint refers to the fact that q&vap, skin and function of surrounding air temperature Tair for some values of mean ′′ radiant temperature Trad . q&evap, resp cannot be positive (occupant may transfer heat to ambient by evapotranspiration but the opposite does not occur). In order to calculate those heat fluxes, values for the saturation pressure P In line with the second approach (i.e., occupant as a control sat,air ′′ were obtained from steam tables as a function of Tair assuming mass), irreversibility rates per unit of area I& were assessed φ ′′ assuming T = T in Eq. (10). Using the values for T and the relative humidity as = 1 (Table 1) and either q&vap, skin or 0 air rad ′′ variation range for Tair as considered in Fig. 2, Fig. 3 shows the q& was deliberately set to zero whenever Psat,air led them to a evap, resp values resulting for the irreversibility rate per unit of area I&′′ . prohibitive value (as discussed above). Except for relatively low air temperatures ( Tair ~ 17°C) where values As temperatures Tair and Trad dictate the heat fluxes yielding the ′′ ′′ + ′′ for I& are comparable to some extent (and among the highest ones), sum q&conv, cl q& clrad, from Eq. (5), another important constraint irreversibility rate levels increase inasmuch as Trad decreases, i.e., as refers to the clothed-body surface temperature Tcl obtained from the Trad departs from Tbody . Bearing in mind the irreversibility rates iterative procedure previously described. As discussed in section 2, obtained for relatively low Trad values, there seems to be a critical the converged value for Tcl is the only one satisfying the ′′ Tair value that minimizes I& . As Trad increases, such critical Tair also homeothermy (i.e., steady-state) assumption of Eq. (5). Still, as far increases and approaches T , as indicated in Table 2. Such as heat transfers to the ambient (= body heat loss) are concerned, body rationale also seems to hold for relatively high Trad (namely, Trad = one should equally bear in mind that any clothing-influenced heat ′′ transfer ultimately starts at the body level so that T > T should 24°C or 26°C), but minimum I& values could not be observed due body cl to the temperature constraints previously discussed in this work and be an additional requisite to allow either q′′ < 0 or q′′ < 0 . &conv, cl & clrad, pointed out in Fig. 2.

454 / Vol. XXXIV, No. 4, October-December 2012 ABCM Human Thermal Comfort: an Irreversibility-Based Approach Emulating Empirical Clothed-Body Correlations…

7.0 evaluated via heat transfer rates, Eqs. (10), (11) and (12), may lead to inaccuracy with respect to the entropy balance. On the other hand, 6.5 ′′ ) one may observe that variations of irreversibility rates I&vap, skin , -2 6.0 ′′ ′′ ′′ ′′ 5.5 I&sw, skin , I&evap, resp , I&conv, resp and I&met (as a function of Tair ) are ′′ ′′ 5.0 sensibly minor when compared to those for I&conv, cl and I& clrad, . 4.5 T_rad = 26°C Respiration and sweat irreversibility rates were also found to be T_rad = 24°C 4.0 negligible in a different calculation approach, as pointed by Batato T_rad = 22°C et al. (1990) under the same conditions considered herein, i.e., 3.5 T_rad = 20°C irreversibility(W.m rate human body at rest (sitting person wearing light clothes). Batato et 3.0 T_rad = 18°C al. (1990) suggested another approach to human thermal comfort T_rad = 16°C 2.5 where three different temperature levels are modelled and submitted 17 19 21 23 25 27 29 31 33 to exergetic analysis. Respiration exergy rates also presented minor values in the studies performed by Mady et al. (2012). Tair (°C) Figure 3. Irreversibility rates per unit of (body surface) area " as a Over the same values and range for Trad and Tair , plots in Fig. 4 function of surrounding air temperature Tair for some values of mean present the behaviour of each contribution to the total irreversibility radiant temperature T . rad rate &′′ . One may also verify that, as T increases, so does the air I rad temperature T at which related contributions become equal (i.e., air I&′′ = I& ′′ ). It is worth remembering that the corresponding Table 2. Minimum irreversibility rates " and corresponding minimizing air conv, cl clrad, temperatures Tair for some mean radiant temperatures Trad . heat transfers occur at TERs at distinct temperatures (namely, Tair Radiant temperature Irreversibility rate Air temperature and Trad ) while relying on a common auxiliary temperature Tcl o 2 o determined from an iterative procedure satisfying a steady-state T ( C) " (W/m ) T ( C) rad air energy balance (i.e., the conceptual heat balance equation). 16.0 5.601 26.0 Furthermore, by analysing I&′′ as a function solely of T (over 18.0 4.994 28.0 clrad, rad ≤ ≤ ′′ 20.0 4.383 29.5 the range 17°C Tair 33°C), it is worth noting that I& clrad, reduces 22.0 3.773 31.5 as Trad increases and approaches to Tbody , which may suggest a link between thermal comfort and exergy loss reduction as far as such Irreversibilities reduce thermodynamic efficiency and they can heat transfer mechanism is concerned. be either intrinsic or avoidable (Kotas, 1985). The latter could be associated to behaviour-influenced irreversibilities (clothing and activity), while the former could be linked to irreversibilities related 100 to basal metabolism or, instead, to some thermal comfort zone. In T_rad = 16°C order to fulfil homeothermy necessities, one could claim that exergy 80 ) deficits should be somehow compensated (e.g., via a shortfall or -2 60 collapse of a thermoregulatory mechanism in detriment of another). 40 Bearing in mind the prospective definition of an irreversibility- 20 based thermal comfort index, those irreversibility dissimilarities in 0 nature could be accounted for. (a) -20 In order to compare each contribution to the total irreversibility I_rad,cl I_conv,cl -40 I_vap,skin I_sw,skin rate per unit of area, one may consider I&′′ in Eq. (10) as the sum of

irreversibilityrate(W.m -60 I_evap,resp I_conv,resp the following irreversibility terms (or components): -80 I_met

′′ = ′′ + ′′ + ′′ + ′′ + -100 I& I&vap, skin I&sw, skin I&evap, resp I&conv, resp 17 19 21 23 25 27 29 31 33 (11) + ′′ + ′′ + ′′ Tair (°C) I&conv, cl I& clrad, I&met 100 T_rad = 18°C where: 80

)

-2 60 T T ′′ −= 0 ′′ ′′ −= 0 ′′ 40 I&vap, skin q&vap, skin , I&sw, skin q&sw, skin , Tair Tair 20 ′′ −= T0 ′′ ′′ −= T0 ′′ I&evap, resp q&evap, resp , I&conv, resp q&conv, resp , 0 Tair Tair (b) -20 (12) I_rad,cl I_conv,cl ′′ −= T0 ′′ ′′ −= T0 ′′ I&conv, cl q&conv, cl , I& clrad, q& clrad, , -40 I_vap,skin I_sw,skin Tair Trad irreversibility (W.m rate -60 I_evap,resp I_conv,resp T ′′ −= 0 ′′ I_met I&met q&met -80 Tbody -100 17 19 21 23 25 27 29 31 33 It is worth noting that Prek´s approach (Prek, 2004, 2005, 2006; Tair (°C) ′′ ′′ ′′ Prek and Butala, 2010), i.e., I&vap, skin , I&sw, skin and I&evap, resp

J. of the Braz. Soc. of Mech. Sci. & Eng. Copyright  2012 by ABCM October-December 2012, Vol. XXXIV, No. 4 / 455 Rabi et al.

100 87 T_rad = 20°C I"_rate conv,cl + I"_rate rad,cl 80 85 ) ) -2 -2 60 83 Trad = 26°C 40 81 Trad = 24°C 20 Trad = 22°C 79 0 Trad = 20°C 77 (c) -20 Trad = 18°C I_rad,cl I_conv,cl 75 -40 I_vap,skin I_sw,skin Trad = 16°C

irreversibility rate (W.m -60 I_evap,resp I_conv,resp 73 irreversibility(W.m rate -80 I_met 71 -100 69 17 19 21 23 25 27 29 31 33 17 19 21 23 25 27 29 31 33 T (°C) air T (°C) air 100 Figure 5. Variation of I&′′ + I& ′′ (clothing-influenced irreversibility T_rad = 22°C conv, cl clrad, 80 rates per unit of area) as a function of air temperature Tair for some values

) of mean radiant temperature Trad .

-2 60 40 ′′ 20 Although temperature Tair influences q&conv, cl while temperature 0 ′′ Trad dictates q& clrad, , it is worth recalling that those clothing- (d) -20 I_rad,cl I_conv,cl influenced heat transfer rates are linked to each other via the -40 I_vap,skin I_sw,skin clothed-body surface temperature Tcl . As already described, the

irreversibilityrate (W.m ′′ + ′′ -60 I_evap,resp I_conv,resp evaluation of the latter depends on the sum q&conv, cl q& clrad, as given -80 I_met by Eq. (5), so that the decrease of one heat transfer mechanism is -100 somehow compensated by the augmentation of the other. In view of 17 19 21 23 25 27 29 31 33 ′′ + ′′ that, Fig. 5 shows the behaviour of the sum I&conv, cl I& clrad, over the Tair (°C) same values and range for Trad and Tair . 100 ′′ T_rad = 24°C Similarly to the behaviour of I& clrad, rates, one can observe that 80 ′′ + ′′ ) the sum I&conv, cl I& clrad, reduces as Trad approaches Tbody . One also

-2 60 ′′ + ′′ 40 verifies that I&conv, cl I& clrad, diminishes inasmuch as Tair decreases 20 ′′ ′′ and, hence, inasmuch as q&conv, cl increases and compensates q& clrad, 0 more and more. In other words, exergy (useful energy) losses are (e) -20 I_rad,cl I_conv,cl minor whenever the body is able to transfer more heat via -40 I_vap,skin I_sw,skin convection.

irreversibility (W.m rate -60 I_evap,resp I_conv,resp For thermoregulation purposes, those previous results may point -80 I_met to a prospective body’s “preference” for losing heat via convection -100 rather than by thermal radiation. Indeed, bearing in mind Eqs. (6c) ′′ 17 19 21 23 25 27 29 31 33 and (6d), by increasing Trad one increases I& clrad, while decreasing T (°C) air ′′ ′′ I&conv, cl . In view of Eq. (6f), I&conv, resp also decreases, though to 100 T_rad = 26°C minor extent if compared to other irreversibility rates, as pointed out 80 previously. )

-2 60 As far as thermoregulation is concerned, there has been some 40 dispute about what variables are in effect the regulated ones from 20 the control theory viewpoint (Blight, 1985) and the “nominees” have comprised distinct temperatures (e.g., core, skin, or brain), 0 body energy content, or heat outflow rate. While widely accepted (f) -20 I_rad,cl I_conv,cl indices have been proposed for uniform thermal ambient, it is -40 I_vap,skin I_sw,skin claimed that thermally non-uniform surroundings still lacks a

irreversibility (W.m rate -60 I_evap,resp I_conv,resp universal index (Zhang and Zhao, 2008). By taking into account -80 I_met thermodynamic factors related to both ambient and the occupant, -100 exergy loss may become another promising “contender” in the 17 19 21 23 25 27 29 31 33 previous list of thermoregulated variables. Indeed, exergy is known

Tair (°C) to increase as the thermodynamic state of a system departs from that of its surroundings so that exergy has the ability to evenly assess Figure 4. Components of the irreversibility rate per unit of area as a function of Tair for: (a) Trad = 16°C, (b) Trad = 18°C, (c) Trad = 20°C, (d) Trad = both extremes of thermal comfort or sensation scale, spanning from 22°C, (e) Trad = 24°C and (f) Trad = 26°C. the uncomfortably cold up to the uncomfortably hot. Results in the literature have indicated that humans perceive hot

or cold discomfort more intensively as thermal conditions depart

456 / Vol. XXXIV, No. 4, October-December 2012 ABCM Human Thermal Comfort: an Irreversibility-Based Approach Emulating Empirical Clothed-Body Correlations… more and more from a given comfort range (Parsons, 1993; losses become lower whenever the body is able to transfer more heat ASHRAE, 2001). Recalling that exergy increases as the via convection other than by thermal radiation, which could be thermodynamic state of any system departs from that of its linked to a prospective “preference” for losing heat via convection surroundings and expecting a similar trend for irreversibilities for thermoregulation purposes. (exergy losses), one could wonder whether thermoregulation has Despite results may eventually encourage the definition of an exergy-saving or entropy-efficient criteria. If so, one may think of irreversibility-based thermal comfort index, it should be stressed irreversibility-based indices as prospective quantitative parameters that irreversibilities were here calculated relying on a steady-state to assess thermal comfort, based on a rational approach combining assumption, which can be quite restrictive. Including other both first and second laws of thermodynamics. scenarios, potential links to thermal comfort claim for future Last but not least, it is worth remembering that steady-state was analysis or improvement as, for example, eventual links between assumed for preliminary calculations herein performed, the analysis intrinsic irreversibilities and basal metabolic and/or comfort zone of transient equations is beyond the scope of this work. Yet, it is levels as well as between avoidable irreversibilities and the known that beyond a critical head temperature (set-point) a sharp departure of such minimum physiological condition. shift occurs from heat loss through insensible evaporation to heat Bearing in mind second-law efficiency of thermoregulatory loss via sweating (Guyton, 1995). With respect to the metabolic heat mechanisms for a resting person, this paper attempted to put forward release, there is also a similar set-point shift from basal to shivering- a model framework to identify minimum irreversibility rates for a induced and it is interesting to observe that both set-points depend wide range of air temperature. As it is well known, minimum on skin temperature. Reminding that humans are able to either irreversibility rates imply minimum entropy generation and, thus, acclimatize or acclimate, i.e., to naturally or artificially acquire maximum second-law efficiency. Finally, as the literature reviewed physiological response changes after long exposure to either hot or in this work suggests, comfort thermal sensation is clearly attached cold environments, it is worth investigating potential connections to minimum human-body exergy consumption-rates, which is also between the coefficient H introduced by OTCI definition expected to render neutral thermal sensation vote (TSV) or neutral (Boregowda, Tiwari and Chaturvedi, 2001) and the basic factors for predicted mean vote (PMV). thermal comfort. One should recall that OTCI depends on entropy Future developments may improve or extend the proposed generation rates while calculations herein presented depend on model framework as, for example, ( i) broader range of Trad values, irreversibility rates ( I& ) or irreversibility rates per unit of area ( I&′′ ). (ii ) distinct body conditions (e.g., performing physical activity) and (iii ) different boundary conditions (e.g., surroundings or climatic However, by properly identifying a reference temperature T0 in the surroundings, one can relate these irreversibility rates by means of regions). It is believed that a comprehensive exergy balance of the Gouy-Stodola relation (Kotas, 1985; Bejan, 1988; Szargut, Morris human body may lead to a better understanding of human thermal and Steward, 1988). comfort and its relation to irreversibility rates.

Concluding Remarks References As suggested elsewhere (e.g., definition of PMV - predicted ASHRAE - American Society of Heating, Refrigeration and Air- Conditioning Engineers, 2001, “Fundamentals Handbook”, Atlanta, USA, mean vote and PPD - predicted percentage dissatisfied), energy Chapter 8. balance (i.e., first-law analysis) for occupant’s body is a necessary Batato, M., Borel, L., Deriaz, O. And Jequier, E., 1990, “Analyse condition for thermal comfort but it may not be a sufficient one. exergétique, théorique et expérimentale du corps humain”, Entropie , Vol. 26, Recognizing that both ambient and personal (human) conditions pp. 120-130. affect thermal sensations, one may attempt to assess either hotness Bejan, A., 1988, “Advanced Engineering Thermodynamics”, John or coldness experiences by taking into account the second-law of Wiley & Sons, New York, USA. thermodynamics as well. By lumping information related to both Bligh, J., 1985, “Regulation of body temperature in man and other ambient and occupant, an analysis combining first and second laws mammals”, in: Shitzer, A. and Eberhart, R.C. (Eds.), “Heat Transfer in Medicine and Biology”, Plenum Press, New York, USA. may enable one to identify process thermodynamic inefficiencies as Boregowda, S.C., Tiwari, S.N. and Chaturvedi, S.K., 2001, “Entropy exergy losses (irreversibilities) thus helping one to quantify thermal generation method to quantify thermal comfort”, Human in efficiency either at optimal conditions (minimum exergy loss) or at Extreme Environments , Vol. 6, pp. 40-45. non-optimal conditions, with thermoregulatory mechanisms Fanger, P.O., 1970, “Thermal Comfort”, McGraw-Hill, New York, USA. rendering thermal discomfort. Gagge, A.P., Stolwijk, J.A.J., and Nishi, Y., 1971, “An effective An underlying question refers to the human body susceptibility temperature scale based on a simple model of human physiological to irreversibilities resulting from its own thermoregulatory regulatory response”, ASHRAE Transactions , Vol. 77, pp. 247-262. mechanisms and this paper suggested more insights to exergetic Givoni, B., 1981, “Man, Climate and Architecture”, Applied Science Publishers, London, UK. analysis of thermal comfort, bearing in mind a prospective Guyton, A.C., 1995, “Textbook of Medical ”, W.B. definition of an irreversibility-based index. Accordingly, Saunders, Philadelphia, USA. preliminary calculations were performed following a model Humphreys, M.A. and Hancock, M., 2007, “Do people like to feel framework that evoked empirical heat-transfer correlations for a ‘neutral’? Exploring the variation of the desired thermal sensation on the human clothed body and the so-called conceptual heat balance ASHRAE scale”, Energy and Buildings , Vol. 39, pp. 867-874. equation, which is based on a steady-state energy balance for the Ivanov, K.P., 2006, “The development of the concepts of homeothermy human body. and thermoregulation”, Journal of Thermal Biology , Vol. 31, pp. 24-29. For a sitting person under distinct combinations of ambient air Kotas, T.J., 1985, “The Exergy Method of Thermal Plant Analysis”, Butterworths, London, UK. and mean radiant temperatures, results showed that a minimum Mady, C.E.K., Ferreira, M.S., Yanagihara, J.I., Saldiva, P.H.N. and irreversibility rate may exist for particular combinations of ambient Oliveira Junior, S., 2012, “Modeling the exergy behavior of human body”, air and mean radiant temperatures. In addition, by comparing Energy , article in press, 8 p. irreversibility contributions from each energy interaction, results Nishi, Y., 1981, “Measurement of thermal balance of man”, in: Cena, K. suggested a stronger influence of both clothing-influenced heat and Clark, J.A. (Eds.), “Bioengineering Thermal Physiology and Comfort”, transfers’ mechanisms, i.e., convective and radiative heat transfers, Elsevier, Amsterdam, The Netherlands. on total irreversibility rate. Among those two mechanisms, exergy

J. of the Braz. Soc. of Mech. Sci. & Eng. Copyright  2012 by ABCM October-December 2012, Vol. XXXIV, No. 4 / 457 Rabi et al.

Olesen, B.W., 1982, “Thermal Comfort”, Technical Review, No. 2, considerations and prospective extensions”, Proceedings of the 11 th Brazilian Bruel & Kjær, Nærum, Denmark. Congress of Thermal Sciences and Engineering, Curitiba, Brazil, Paper Parsons, K.C., 1993, “Human Thermal Environments”, Taylor & CIT06-0497, CD-ROM. Francis, London, UK. Simone, A., Kolarik, J., Iwamatsu, T., Asada, H., Dovjak, M., Schellen, Prek, M., 2004, “Exergy analysis of thermal comfort”, International L., Shukuya, M. and Olesen, B.W., 2011, “A relation between calculated Journal of Exergy , Vol. 1, pp. 303-315. human body exergy consumption rate and subjectively assessed thermal Prek, M., 2005, “Thermodynamic analysis of human heat and mass sensation”, Energy and Buildings , Vol. 43, pp. 1-9. transfer and their impact on thermal comfort”, International Journal of Heat Szargut, J., Morris, D.R. and Steward, F.R., 1988, “Exergy Analysis of and Mass Transfer , Vol. 48, pp. 731-739. Thermal, Chemical and Metallurgical Process”, Hemisphere, New York, USA. Prek, M., 2006, “Thermodynamical analysis of human thermal comfort”, Webb, P., 1995, “The physiology of heat regulation”, American Journal Energy , Vol. 31, pp. 732-743. of Physiology - Regulatory, Integrative and Comparative Physiology , Vol. Prek, M. and Butala, V., 2010, “Principles of exergy analysis of human 268, pp. R838-R850. heat and mass exchange with the indoor environment”, International Journal Zhang, Y. and Zhao, R., 2008, “Overall thermal sensation, acceptability of Heat and Mass Transfer , Vol. 53, pp. 5806-5814. and comfort”, Building and Environment , Vol. 43, pp. 44-50. Rabi, J.A., Fresia, C.E.S. and Benzal, G., 2006, “Thermal comfort and second-law analysis of thermoregulation mechanisms: preliminary

458 / Vol. XXXIV, No. 4, October-December 2012 ABCM