2.5. Momentum, Sensible Heat, and Latent Heat Fluxes

The zonal \tau _{x} and meridional \tau _{y} momentum fluxes (kg m-1 s-2), sensible heat flux H (W m-2), and water vapor flux E (kg m-2 s-1) between the atmosphere at reference height z_{atm,\, x} (m) [where x is height for wind (momentum) (m), temperature (sensible heat) (h), and humidity (water vapor) (w); with zonal and meridional winds u_{atm} and v_{atm} (m s-1), potential temperature \theta _{atm} (K), and specific humidity q_{atm} (kg kg-1)] and the surface [with u_{s} , v_{s} , \theta _{s} , and q_{s} ] are

(2.5.1)\tau _{x} =-\rho _{atm} \frac{\left(u_{atm} -u_{s} \right)}{r_{am} }

(2.5.2)\tau _{y} =-\rho _{atm} \frac{\left(v_{atm} -v_{s} \right)}{r_{am} }

(2.5.3)H=-\rho _{atm} C_{p} \frac{\left(\theta _{atm} -\theta _{s} \right)}{r_{ah} }

(2.5.4)E=-\rho _{atm} \frac{\left(q_{atm} -q_{s} \right)}{r_{aw} } .

These fluxes are derived in the next section from Monin-Obukhov similarity theory developed for the surface layer (i.e., the nearly constant flux layer above the surface sublayer). In this derivation, u_{s} and v_{s} are defined to equal zero at height z_{0m} +d (the apparent sink for momentum) so that r_{am} is the aerodynamic resistance (s m-1) for momentum between the atmosphere at height z_{atm,\, m} and the surface at height z_{0m} +d. Thus, the momentum fluxes become

(2.5.5)\tau _{x} =-\rho _{atm} \frac{u_{atm} }{r_{am} }

(2.5.6)\tau _{y} =-\rho _{atm} \frac{v_{atm} }{r_{am} } .

Likewise, \theta _{s} and q_{s} are defined at heights z_{0h} +d and z_{0w} +d (the apparent sinks for heat and water vapor, respectively

r_{aw} are the aerodynamic resistances (s m-1) to sensible heat and water vapor transfer between the atmosphere at heights z_{atm,\, h} and z_{atm,\, w} and the surface at heights z_{0h} +d and z_{0w} +d, respectively. The specific heat capacity of air C_{p} (J kg-1 K-1) is a constant (Table 2.2.7). The atmospheric potential temperature used here is

(2.5.7)\theta _{atm} =T_{atm} +\Gamma _{d} z_{atm,\, h}

where T_{atm} is the air temperature (K) at height z_{atm,\, h} and \Gamma _{d} =0.0098 K m-1 is the negative of the dry adiabatic lapse rate [this expression is first-order equivalent to \theta _{atm} =T_{atm} \left({P_{srf} \mathord{\left/ {\vphantom {P_{srf}  P_{atm} }} \right. \kern-\nulldelimiterspace} P_{atm} } \right)^{{R_{da} \mathord{\left/ {\vphantom {R_{da}  C_{p} }} \right. \kern-\nulldelimiterspace} C_{p} } } (Stull 1988), where P_{srf} is the surface pressure (Pa), P_{atm} is the atmospheric pressure (Pa), and R_{da} is the gas constant for dry air (J kg-1 K-1) (Table 2.2.7)]. By definition, \theta _{s} =T_{s} . The density of moist air (kg m-3) is

(2.5.8)\rho _{atm} =\frac{P_{atm} -0.378e_{atm} }{R_{da} T_{atm} }

where the atmospheric vapor pressure e_{atm} (Pa) is derived from the atmospheric specific humidity q_{atm}

(2.5.9)e_{atm} =\frac{q_{atm} P_{atm} }{0.622+0.378q_{atm} } .

2.5.1. Monin-Obukhov Similarity Theory

The surface vertical kinematic fluxes of momentum \overline{u'w'} and \overline{v'w'} (m2 s-2), sensible heat \overline{\theta 'w'} (K m s -1), and latent heat \overline{q'w'} (kg kg-1 m s-1), where u', v', w', \theta ', and q' are zonal horizontal wind, meridional horizontal wind, vertical velocity, potential temperature, and specific humidity turbulent fluctuations about the mean, are defined from Monin-Obukhov similarity applied to the surface layer. This theory states that when scaled appropriately, the dimensionless mean horizontal wind speed, mean potential temperature, and mean specific humidity profile gradients depend on unique functions of \zeta =\frac{z-d}{L} (Zeng et al. 1998) as

(2.5.10)\frac{k\left(z-d\right)}{u_{*} } \frac{\partial \left|{\it u}\right|}{\partial z} =\phi _{m} \left(\zeta \right)

(2.5.11)\frac{k\left(z-d\right)}{\theta _{*} } \frac{\partial \theta }{\partial z} =\phi _{h} \left(\zeta \right)

(2.5.12)\frac{k\left(z-d\right)}{q_{*} } \frac{\partial q}{\partial z} =\phi _{w} \left(\zeta \right)

where z is height in the surface layer (m), d is the displacement height (m), L is the Monin-Obukhov length scale (m) that accounts for buoyancy effects resulting from vertical density gradients (i.e., the atmospheric stability), k is the von Karman constant (Table 2.2.7), and \left|{\it u}\right| is the atmospheric wind speed (m s-1). \phi _{m} , \phi _{h} , and \phi _{w} are universal (over any surface) similarity functions of \zeta that relate the constant fluxes of momentum, sensible heat, and latent heat to the mean profile gradients of \left|{\it u}\right|, \theta , and q in the surface layer. In neutral conditions, \phi _{m} =\phi _{h} =\phi _{w} =1. The velocity (i.e., friction velocity) u_{\*} (m s-1), temperature \theta _{\*} (K), and moisture q_{\*} (kg kg-1) scales are

(2.5.13)u_{*}^{2} =\sqrt{\left(\overline{u'w'}\right)^{2} +\left(\overline{v'w'}\right)^{2} } =\frac{\left|{\it \tau }\right|}{\rho _{atm} }

(2.5.14)\theta _{*} u_{*} =-\overline{\theta 'w'}=-\frac{H}{\rho _{atm} C_{p} }

(2.5.15)q_{*} u_{*} =-\overline{q'w'}=-\frac{E}{\rho _{atm} }

where \left|{\it \tau }\right| is the shearing stress (kg m-1 s-2), with zonal and meridional components \overline{u'w'}=-\frac{\tau _{x} }{\rho _{atm} } and \overline{v'w'}=-\frac{\tau _{y} }{\rho _{atm} } , respectively, H is the sensible heat flux (W m-2) and E is the water vapor flux (kg m-2 s-1).

The length scale L is the Monin-Obukhov length defined as

(2.5.16)L=-\frac{u_{*}^{3} }{k\left(\frac{g}{\overline{\theta _{v,\, atm} }} \right)\theta '_{v} w'} =\frac{u_{*}^{2} \overline{\theta _{v,\, atm} }}{kg\theta _{v*} }

where g is the acceleration of gravity (m s-2) (Table 2.2.7), and \overline{\theta _{v,\, atm} }=\overline{\theta _{atm} }\left(1+0.61q_{atm} \right) is the reference virtual potential temperature. L>0 indicates stable conditions. L<0 indicates unstable conditions. L=\infty for neutral conditions. The temperature scale \theta _{v*} is defined as

(2.5.17)\theta _{v*} u_{*} =\left[\theta _{*} \left(1+0.61q_{atm} \right)+0.61\overline{\theta _{atm} }q_{*} \right]u_{*}

where \overline{\theta _{atm} } is the atmospheric potential temperature.

Following Panofsky and Dutton (1984), the differential equations for \phi _{m} \left(\zeta \right), \phi _{h} \left(\zeta \right), and \phi _{w} \left(\zeta \right) can be integrated formally without commitment to their exact forms. Integration between two arbitrary heights in the surface layer z_{2} and z_{1} (z_{2} >z_{1} ) with horizontal winds \left|{\it u}\right|_{1} and \left|{\it u}\right|_{2} , potential temperatures \theta _{1} and \theta _{2} , and specific humidities q_{1} and q_{2} results in

(2.5.18)\left|{\it u}\right|_{2} -\left|{\it u}\right|_{1} =\frac{u_{*} }{k} \left[\ln \left(\frac{z_{2} -d}{z_{1} -d} \right)-\psi _{m} \left(\frac{z_{2} -d}{L} \right)+\psi _{m} \left(\frac{z_{1} -d}{L} \right)\right]

(2.5.19)\theta _{2} -\theta _{1} =\frac{\theta _{*} }{k} \left[\ln \left(\frac{z_{2} -d}{z_{1} -d} \right)-\psi _{h} \left(\frac{z_{2} -d}{L} \right)+\psi _{h} \left(\frac{z_{1} -d}{L} \right)\right]

(2.5.20)q_{2} -q_{1} =\frac{q_{*} }{k} \left[\ln \left(\frac{z_{2} -d}{z_{1} -d} \right)-\psi _{w} \left(\frac{z_{2} -d}{L} \right)+\psi _{w} \left(\frac{z_{1} -d}{L} \right)\right].

The functions \psi _{m} \left(\zeta \right), \psi _{h} \left(\zeta \right), and \psi _{w} \left(\zeta \right) are defined as

(2.5.21)\psi _{m} \left(\zeta \right)=\int _{{z_{0m} \mathord{\left/ {\vphantom {z_{0m}  L}} \right. \kern-\nulldelimiterspace} L} }^{\zeta }\frac{\left[1-\phi _{m} \left(x\right)\right]}{x} \, dx

(2.5.22)\psi _{h} \left(\zeta \right)=\int _{{z_{0h} \mathord{\left/ {\vphantom {z_{0h}  L}} \right. \kern-\nulldelimiterspace} L} }^{\zeta }\frac{\left[1-\phi _{h} \left(x\right)\right]}{x} \, dx

(2.5.23)\psi _{w} \left(\zeta \right)=\int _{{z_{0w} \mathord{\left/ {\vphantom {z_{0w}  L}} \right. \kern-\nulldelimiterspace} L} }^{\zeta }\frac{\left[1-\phi _{w} \left(x\right)\right]}{x} \, dx

where z_{0m} , z_{0h} , and z_{0w} are the roughness lengths (m) for momentum, sensible heat, and water vapor, respectively.

Defining the surface values

\left|{\it u}\right|_{1} =0{\rm \; at\; }z_{1} =z_{0m} +d,

\theta _{1} =\theta _{s} {\rm \; at\; }z_{1} =z_{0h} +d,{\rm \; and}

q_{1} =q_{s} {\rm \; at\; }z_{1} =z_{0w} +d,

and the atmospheric values at z_{2} =z_{atm,\, x}

(2.5.24)\left|{\it u}\right|_{2} =V_{a} {\rm =\; }\sqrt{u_{atm}^{2} +v_{atm}^{2} +U_{c}^{2} } \ge 1,

\theta _{2} =\theta _{atm} {\rm ,\; and}

q_{2} =q_{atm} {\rm ,\; }

the integral forms of the flux-gradient relations are

(2.5.25)V_{a} =\frac{u_{*} }{k} \left[\ln \left(\frac{z_{atm,\, m} -d}{z_{0m} } \right)-\psi _{m} \left(\frac{z_{atm,\, m} -d}{L} \right)+\psi _{m} \left(\frac{z_{0m} }{L} \right)\right]

(2.5.26)\theta _{atm} -\theta _{s} =\frac{\theta _{*} }{k} \left[\ln \left(\frac{z_{atm,\, h} -d}{z_{0h} } \right)-\psi _{h} \left(\frac{z_{atm,\, h} -d}{L} \right)+\psi _{h} \left(\frac{z_{0h} }{L} \right)\right]

(2.5.27)q_{atm} -q_{s} =\frac{q_{*} }{k} \left[\ln \left(\frac{z_{atm,\, w} -d}{z_{0w} } \right)-\psi _{w} \left(\frac{z_{atm,\, w} -d}{L} \right)+\psi _{w} \left(\frac{z_{0w} }{L} \right)\right].

The constraint V_{a} \ge 1 is required simply for numerical reasons to prevent H and E from becoming small with small wind speeds. The convective velocity U_{c} accounts for the contribution of large eddies in the convective boundary layer to surface fluxes as follows

(2.5.28)U_{c} = \left\{
\begin{array}{ll}
0 & \qquad \zeta \ge {\rm 0} \quad {\rm (stable)} \\
\beta w_{*} & \qquad \zeta < 0 \quad {\rm (unstable)}
\end{array} \right\}

where w_{*} is the convective velocity scale

(2.5.29)w_{*} =\left(\frac{-gu_{\*} \theta _{v*} z_{i} }{\overline{\theta _{v,\, atm} }} \right)^{{1\mathord{\left/ {\vphantom {1 3}} \right. \kern-\nulldelimiterspace} 3} } ,

z_{i} =1000 is the convective boundary layer height (m), and \beta =1.

The momentum flux gradient relations are (Zeng et al. 1998)

(2.5.30)\begin{array}{llr}
\phi _{m} \left(\zeta \right)=0.7k^{{2\mathord{\left/ {\vphantom {2 3}} \right. \kern-\nulldelimiterspace} 3} } \left(-\zeta \right)^{{1\mathord{\left/ {\vphantom {1 3}} \right. \kern-\nulldelimiterspace} 3} } & \qquad {\rm for\; }\zeta <-1.574 & \ {\rm \; (very\; unstable)} \\
\phi _{m} \left(\zeta \right)=\left(1-16\zeta \right)^{-{1\mathord{\left/ {\vphantom {1 4}} \right. \kern-\nulldelimiterspace} 4} } & \qquad {\rm for\; -1.574}\le \zeta <0 & \ {\rm \; (unstable)} \\
\phi _{m} \left(\zeta \right)=1+5\zeta & \qquad {\rm for\; }0\le \zeta \le 1& \ {\rm \; (stable)} \\
\phi _{m} \left(\zeta \right)=5+\zeta & \qquad {\rm for\; }\zeta  >1 & \ {\rm\; (very\; stable).}
\end{array}

The sensible and latent heat flux gradient relations are (Zeng et al. 1998)

(2.5.31)\begin{array}{llr}
\phi _{h} \left(\zeta \right)=\phi _{w} \left(\zeta \right)=0.9k^{{4\mathord{\left/ {\vphantom {4 3}} \right. \kern-\nulldelimiterspace} 3} } \left(-\zeta \right)^{{-1\mathord{\left/ {\vphantom {-1 3}} \right. \kern-\nulldelimiterspace} 3} } & \qquad {\rm for\; }\zeta <-0.465 & \ {\rm \; (very\; unstable)} \\
\phi _{h} \left(\zeta \right)=\phi _{w} \left(\zeta \right)=\left(1-16\zeta \right)^{-{1\mathord{\left/ {\vphantom {1 2}} \right. \kern-\nulldelimiterspace} 2} } & \qquad {\rm for\; -0.465}\le \zeta <0 & \ {\rm \; (unstable)} \\
\phi _{h} \left(\zeta \right)=\phi _{w} \left(\zeta \right)=1+5\zeta & \qquad {\rm for\; }0\le \zeta \le 1 & \ {\rm \; (stable)} \\
\phi _{h} \left(\zeta \right)=\phi _{w} \left(\zeta \right)=5+\zeta & \qquad {\rm for\; }\zeta  >1 & \ {\rm \; (very\; stable).}
\end{array}

To ensure continuous functions of \phi _{m} \left(\zeta \right), \phi _{h} \left(\zeta \right), and \phi _{w} \left(\zeta \right), the simplest approach (i.e., without considering any transition regimes) is to match the relations for very unstable and unstable conditions at \zeta _{m} =-1.574 for \phi _{m} \left(\zeta \right) and \zeta _{h} =\zeta _{w} =-0.465 for \phi _{h} \left(\zeta \right)=\phi _{w} \left(\zeta \right) (Zeng et al. 1998). The flux gradient relations can be integrated to yield wind profiles for the following conditions:

Very unstable \left(\zeta <-1.574\right)

(2.5.32)V_{a} =\frac{u_{*} }{k} \left\{\left[\ln \frac{\zeta _{m} L}{z_{0m} } -\psi _{m} \left(\zeta _{m} \right)\right]+1.14\left[\left(-\zeta \right)^{{1\mathord{\left/ {\vphantom {1 3}} \right. \kern-\nulldelimiterspace} 3} } -\left(-\zeta _{m} \right)^{{1\mathord{\left/ {\vphantom {1 3}} \right. \kern-\nulldelimiterspace} 3} } \right]+\psi _{m} \left(\frac{z_{0m} }{L} \right)\right\}

Unstable \left(-1.574\le \zeta <0\right)

(2.5.33)V_{a} =\frac{u_{*} }{k} \left\{\left[\ln \frac{z_{atm,\, m} -d}{z_{0m} } -\psi _{m} \left(\zeta \right)\right]+\psi _{m} \left(\frac{z_{0m} }{L} \right)\right\}

Stable \left(0\le \zeta \le 1\right)

(2.5.34)V_{a} =\frac{u_{*} }{k} \left\{\left[\ln \frac{z_{atm,\, m} -d}{z_{0m} } +5\zeta \right]-5\frac{z_{0m} }{L} \right\}

Very stable \left(\zeta >1\right)

(2.5.35)V_{a} =\frac{u_{*} }{k} \left\{\left[\ln \frac{L}{z_{0m} } +5\right]+\left[5\ln \zeta +\zeta -1\right]-5\frac{z_{0m} }{L} \right\}

where

(2.5.36)\psi _{m} \left(\zeta \right)=2\ln \left(\frac{1+x}{2} \right)+\ln \left(\frac{1+x^{2} }{2} \right)-2\tan ^{-1} x+\frac{\pi }{2}

and

x=\left(1-16\zeta \right)^{{1\mathord{\left/ {\vphantom {1 4}} \right. \kern-\nulldelimiterspace} 4} } .

The potential temperature profiles are:

Very unstable \left(\zeta <-0.465\right)

(2.5.37)\theta _{atm} -\theta _{s} =\frac{\theta _{*} }{k} \left\{\left[\ln \frac{\zeta _{h} L}{z_{0h} } -\psi _{h} \left(\zeta _{h} \right)\right]+0.8\left[\left(-\zeta _{h} \right)^{{-1\mathord{\left/ {\vphantom {-1 3}} \right. \kern-\nulldelimiterspace} 3} } -\left(-\zeta \right)^{{-1\mathord{\left/ {\vphantom {-1 3}} \right. \kern-\nulldelimiterspace} 3} } \right]+\psi _{h} \left(\frac{z_{0h} }{L} \right)\right\}

Unstable \left(-0.465\le \zeta <0\right)

(2.5.38)\theta _{atm} -\theta _{s} =\frac{\theta _{*} }{k} \left\{\left[\ln \frac{z_{atm,\, h} -d}{z_{0h} } -\psi _{h} \left(\zeta \right)\right]+\psi _{h} \left(\frac{z_{0h} }{L} \right)\right\}

Stable \left(0\le \zeta \le 1\right)

(2.5.39)\theta _{atm} -\theta _{s} =\frac{\theta _{*} }{k} \left\{\left[\ln \frac{z_{atm,\, h} -d}{z_{0h} } +5\zeta \right]-5\frac{z_{0h} }{L} \right\}

Very stable \left(\zeta >1\right)

(2.5.40)\theta _{atm} -\theta _{s} =\frac{\theta _{*} }{k} \left\{\left[\ln \frac{L}{z_{0h} } +5\right]+\left[5\ln \zeta +\zeta -1\right]-5\frac{z_{0h} }{L} \right\}.

The specific humidity profiles are:

Very unstable \left(\zeta <-0.465\right)

(2.5.41)q_{atm} -q_{s} =\frac{q_{*} }{k} \left\{\left[\ln \frac{\zeta _{w} L}{z_{0w} } -\psi _{w} \left(\zeta _{w} \right)\right]+0.8\left[\left(-\zeta _{w} \right)^{{-1\mathord{\left/ {\vphantom {-1 3}} \right. \kern-\nulldelimiterspace} 3} } -\left(-\zeta \right)^{{-1\mathord{\left/ {\vphantom {-1 3}} \right. \kern-\nulldelimiterspace} 3} } \right]+\psi _{w} \left(\frac{z_{0w} }{L} \right)\right\}

Unstable \left(-0.465\le \zeta <0\right)

(2.5.42)q_{atm} -q_{s} =\frac{q_{*} }{k} \left\{\left[\ln \frac{z_{atm,\, w} -d}{z_{0w} } -\psi _{w} \left(\zeta \right)\right]+\psi _{w} \left(\frac{z_{0w} }{L} \right)\right\}

Stable \left(0\le \zeta \le 1\right)

(2.5.43)q_{atm} -q_{s} =\frac{q_{*} }{k} \left\{\left[\ln \frac{z_{atm,\, w} -d}{z_{0w} } +5\zeta \right]-5\frac{z_{0w} }{L} \right\}

Very stable \left(\zeta >1\right)

(2.5.44)q_{atm} -q_{s} =\frac{q_{*} }{k} \left\{\left[\ln \frac{L}{z_{0w} } +5\right]+\left[5\ln \zeta +\zeta -1\right]-5\frac{z_{0w} }{L} \right\}

where

(2.5.45)\psi _{h} \left(\zeta \right)=\psi _{w} \left(\zeta \right)=2\ln \left(\frac{1+x^{2} }{2} \right).

Using the definitions of u_{*} , \theta _{*} , and q_{*} , an iterative solution of these equations can be used to calculate the surface momentum, sensible heat, and water vapor flux using atmospheric and surface values for \left|{\it u}\right|, \theta , and q except that L depends on u_{*} , \theta _{*} , and q_{*} . However, the bulk Richardson number

(2.5.46)R_{iB} =\frac{\theta _{v,\, atm} -\theta _{v,\, s} }{\overline{\theta _{v,\, atm} }} \frac{g\left(z_{atm,\, m} -d\right)}{V_{a}^{2} }

is related to \zeta (Arya 2001) as

(2.5.47)R_{iB} =\zeta \left[\ln \left(\frac{z_{atm,\, h} -d}{z_{0h} } \right)-\psi _{h} \left(\zeta \right)\right]\left[\ln \left(\frac{z_{atm,\, m} -d}{z_{0m} } \right)-\psi _{m} \left(\zeta \right)\right]^{-2} .

Using \phi _{h} =\phi _{m}^{2} =\left(1-16\zeta \right)^{-{1\mathord{\left/ {\vphantom {1 2}} \right. \kern-\nulldelimiterspace} 2} } for unstable conditions and \phi _{h} =\phi _{m} =1+5\zeta for stable conditions to determine \psi _{m} \left(\zeta \right) and \psi _{h} \left(\zeta \right), the inverse relationship \zeta =f\left(R_{iB} \right) can be solved to obtain a first guess for \zeta and thus L from

(2.5.48)\begin{array}{lcr}
\zeta =\frac{R_{iB} \ln \left(\frac{z_{atm,\, m} -d}{z_{0m} } \right)}{1-5\min \left(R_{iB} ,0.19\right)} & \qquad 0.01\le \zeta \le 2 & \qquad {\rm for\; }R_{iB} \ge 0 {\rm \; (neutral\; or\; stable)} \\
\zeta =R_{iB} \ln \left(\frac{z_{atm,\, m} -d}{z_{0m} } \right) & \qquad -100\le \zeta \le -0.01 & \qquad {\rm for\; }R_{iB} <0 \ {\rm \; (unstable)}
\end{array}.

Upon iteration (section 2.5.3.2), the following is used to determine \zeta and thus L

(2.5.49)\zeta =\frac{\left(z_{atm,\, m} -d\right)kg\theta _{v*} }{u_{*}^{2} \overline{\theta _{v,\, atm} }}

where

\begin{array}{cr}
0.01\le \zeta \le 2 & \qquad {\rm for\; }\zeta \ge 0{\rm \; (neutral\; or\; stable)} \\
{\rm -100}\le \zeta \le {\rm -0.01} & \qquad {\rm for\; }\zeta <0{\rm \; (unstable)}
\end{array}.

The difference in virtual potential air temperature between the reference height and the surface is

(2.5.50)\theta _{v,\, atm} -\theta _{v,\, s} =\left(\theta _{atm} -\theta _{s} \right)\left(1+0.61q_{atm} \right)+0.61\overline{\theta _{atm} }\left(q_{atm} -q_{s} \right).

The momentum, sensible heat, and water vapor fluxes between the surface and the atmosphere can also be written in the form

(2.5.51)\tau _{x} =-\rho _{atm} \frac{\left(u_{atm} -u_{s} \right)}{r_{am} }

(2.5.52)\tau _{y} =-\rho _{atm} \frac{\left(v_{atm} -v_{s} \right)}{r_{am} }

(2.5.53)H=-\rho _{atm} C_{p} \frac{\left(\theta _{atm} -\theta _{s} \right)}{r_{ah} }

(2.5.54)E=-\rho _{atm} \frac{\left(q_{atm} -q_{s} \right)}{r_{aw} }

where the aerodynamic resistances (s m-1) are

(2.5.55)r_{am} =\frac{V_{a} }{u_{*}^{2} } =\frac{1}{k^{2} V_{a} } \left[\ln \left(\frac{z_{atm,\, m} -d}{z_{0m} } \right)-\psi _{m} \left(\frac{z_{atm,\, m} -d}{L} \right)+\psi _{m} \left(\frac{z_{0m} }{L} \right)\right]^{2}

(2.5.56)\begin{array}{l} {r_{ah} =\frac{\theta _{atm} -\theta _{s} }{\theta _{*} u_{*} } =\frac{1}{k^{2} V_{a} } \left[\ln \left(\frac{z_{atm,\, m} -d}{z_{0m} } \right)-\psi _{m} \left(\frac{z_{atm,\, m} -d}{L} \right)+\psi _{m} \left(\frac{z_{0m} }{L} \right)\right]} \\ {\qquad \left[\ln \left(\frac{z_{atm,\, h} -d}{z_{0h} } \right)-\psi _{h} \left(\frac{z_{atm,\, h} -d}{L} \right)+\psi _{h} \left(\frac{z_{0h} }{L} \right)\right]} \end{array}

(2.5.57)\begin{array}{l} {r_{aw} =\frac{q_{atm} -q_{s} }{q_{*} u_{*} } =\frac{1}{k^{2} V_{a} } \left[\ln \left(\frac{z_{atm,\, m} -d}{z_{0m} } \right)-\psi _{m} \left(\frac{z_{atm,\, m} -d}{L} \right)+\psi _{m} \left(\frac{z_{0m} }{L} \right)\right]} \\ {\qquad \left[\ln \left(\frac{z_{atm,\, {\it w}} -d}{z_{0w} } \right)-\psi _{w} \left(\frac{z_{atm,\, w} -d}{L} \right)+\psi _{w} \left(\frac{z_{0w} }{L} \right)\right]} \end{array}.

A 2-m height “screen” temperature is useful for comparison with observations

(2.5.58)T_{2m} =\theta _{s} +\frac{\theta _{*} }{k} \left[\ln \left(\frac{2+z_{0h} }{z_{0h} } \right)-\psi _{h} \left(\frac{2+z_{0h} }{L} \right)+\psi _{h} \left(\frac{z_{0h} }{L} \right)\right]

where for convenience, “2-m” is defined as 2 m above the apparent sink for sensible heat (z_{0h} +d). Similarly, a 2-m height specific humidity is defined as

(2.5.59)q_{2m} =q_{s} +\frac{q_{*} }{k} \left[\ln \left(\frac{2+z_{0w} }{z_{0w} } \right)-\psi _{w} \left(\frac{2+z_{0w} }{L} \right)+\psi _{w} \left(\frac{z_{0w} }{L} \right)\right].

Relative humidity is

(2.5.60)RH_{2m} =\min \left(100,\, \frac{q_{2m} }{q_{sat}^{T_{2m} } } \times 100\right)

where q_{sat}^{T_{2m} } is the saturated specific humidity at the 2-m temperature T_{2m} (section 2.5.5).

A 10-m wind speed is calculated as (note that this is not consistent with the 10-m wind speed calculated for the dust model as described in Chapter 2.31)

(2.5.61)u_{10m} =\left\{\begin{array}{l} {V_{a} \qquad z_{atm,\, m} \le 10} \\ {V_{a} -\frac{u_{*} }{k} \left[\ln \left(\frac{z_{atm,\, m} -d}{10+z_{0m} } \right)-\psi _{m} \left(\frac{z_{atm,\, m} -d}{L} \right)+\psi _{m} \left(\frac{10+z_{0m} }{L} \right)\right]\qquad z_{atm,\, m} >10} \end{array}\right\}

2.5.2. Sensible and Latent Heat Fluxes for Non-Vegetated Surfaces

Surfaces are considered non-vegetated for the surface flux calculations if leaf plus stem area index L+S<0.05 (section 2.2.1.4). By definition, this includes bare soil and glaciers. The solution for lakes is described in Chapter 2.12. For these surfaces, the surface may be exposed to the atmosphere, snow covered, and/or surface water covered, so that the sensible heat flux H_{g} (W m-2) is, with reference to Figure 2.5.1,

(2.5.62)H_{g} =\left(1-f_{sno} -f_{h2osfc} \right)H_{soil} +f_{sno} H_{snow} +f_{h2osfc} H_{h2osfc}

where \left(1-f_{sno} -f_{h2osfc} \right), f_{sno} , and f_{h2osfc} are the exposed, snow covered, and surface water covered fractions of the grid cell. The individual fluxes based on the temperatures of the soil T_{1} , snow T_{snl+1} , and surface water T_{h2osfc} are

(2.5.63)H_{soil} =-\rho _{atm} C_{p} \frac{\left(\theta _{atm} -T_{1} \right)}{r_{ah} }

(2.5.64)H_{sno} =-\rho _{atm} C_{p} \frac{\left(\theta _{atm} -T_{snl+1} \right)}{r_{ah} }

(2.5.65)H_{h2osfc} =-\rho _{atm} C_{p} \frac{\left(\theta _{atm} -T_{h2osfc} \right)}{r_{ah} }

where \rho _{atm} is the density of atmospheric air (kg m-3), C_{p} is the specific heat capacity of air (J kg-1 K-1) (Table 2.2.7), \theta _{atm} is the atmospheric potential temperature (K), and r_{ah} is the aerodynamic resistance to sensible heat transfer (s m-1).

The water vapor flux E_{g} (kg m-2 s-1) is, with reference to Figure 2.5.2,

(2.5.66)E_{g} =\left(1-f_{sno} -f_{h2osfc} \right)E_{soil} +f_{sno} E_{snow} +f_{h2osfc} E_{h2osfc}

(2.5.67)E_{soil} =-\frac{\rho _{atm} \left(q_{atm} -q_{soil} \right)}{r_{aw} + r_{soil}}

(2.5.68)E_{sno} =-\frac{\rho _{atm} \left(q_{atm} -q_{sno} \right)}{r_{aw} }

(2.5.69)E_{h2osfc} =-\frac{\rho _{atm} \left(q_{atm} -q_{h2osfc} \right)}{r_{aw} }

where q_{atm} is the atmospheric specific humidity (kg kg-1), q_{soil} , q_{sno} , and q_{h2osfc} are the specific humidities (kg kg-1) of the soil, snow, and surface water, respectively, r_{aw} is the aerodynamic resistance to water vapor transfer (s m-1), and r _{soi} is the soil resistance to water vapor transfer (s m-1). The specific humidities of the snow q_{sno} and surface water q_{h2osfc} are assumed to be at the saturation specific humidity of their respective temperatures

(2.5.70)q_{sno} =q_{sat}^{T_{snl+1} }

(2.5.71)q_{h2osfc} =q_{sat}^{T_{h2osfc} }

The specific humidity of the soil surface q_{soil} is assumed to be proportional to the saturation specific humidity

(2.5.72)q_{soil} =\alpha _{soil} q_{sat}^{T_{1} }

where q_{sat}^{T_{1} } is the saturated specific humidity at the soil surface temperature T_{1} (section 2.5.5). The factor \alpha _{soil} is a function of the surface soil water matric potential \psi as in Philip (1957)

(2.5.73)\alpha _{soil} =\exp \left(\frac{\psi _{1} g}{1\times 10^{3} R_{wv} T_{1} } \right)

where R_{wv} is the gas constant for water vapor (J kg-1 K-1) (Table 2.2.7), g is the gravitational acceleration (m s-2) (Table 2.2.7), and \psi _{1} is the soil water matric potential of the top soil layer (mm). The soil water matric potential \psi _{1} is

(2.5.74)\psi _{1} =\psi _{sat,\, 1} s_{1}^{-B_{1} } \ge -1\times 10^{8}

where \psi _{sat,\, 1} is the saturated matric potential (mm) (section 2.7.3.1), B_{1} is the Clapp and Hornberger (1978) parameter (section 2.7.3.1), and s_{1} is the wetness of the top soil layer with respect to saturation. The surface wetness s_{1} is a function of the liquid water and ice content

(2.5.75)s_{1} =\frac{1}{\Delta z_{1} \theta _{sat,\, 1} } \left[\frac{w_{liq,\, 1} }{\rho _{liq} } +\frac{w_{ice,\, 1} }{\rho _{ice} } \right]\qquad 0.01\le s_{1} \le 1.0

where \Delta z_{1} is the thickness of the top soil layer (m), \rho _{liq} and \rho _{ice} are the density of liquid water and ice (kg m-3) (Table 2.2.7), w_{liq,\, 1} and w_{ice,\, 1} are the mass of liquid water and ice of the top soil layer (kg m-2) (Chapter 2.7), and \theta _{sat,\, 1} is the saturated volumetric water content (i.e., porosity) of the top soil layer (mm3 mm-3) (section 2.7.3.1). If q_{sat}^{T_{1} } >q_{atm} and q_{atm} >q_{soil} , then q_{soil} =q_{atm} and \frac{dq_{soil} }{dT} =0. This prevents large increases (decreases) in q_{soil} for small increases (decreases) in soil moisture in very dry soils.

The resistance to water vapor transfer occurring within the soil matrix r_{soil} (s m-1) is

(2.5.76)r_{soil} = \frac{DSL}{D_{v} \tau}

where DSL is the thickness of the dry surface layer (m), D_{v} is the molecular diffusivity of water vapor in air (m2 s-2) and \tau (unitless) describes the tortuosity of the vapor flow paths through the soil matrix (Swenson and Lawrence 2014).

The thickness of the dry surface layer is given by

(2.5.77)DSL =
\begin{array}{lr}
D_{max} \ \frac{\left( \theta_{init} - \theta_{1}\right)}
{\left(\theta_{init} - \theta_{air}\right)} & \qquad \theta_{1} < \theta_{init} \\
0 &  \qquad \theta_{1} \ge \theta_{init}
\end{array}

where D_{max} is a parameter specifying the length scale of the maximum DSL thickness (default value = 15 mm), \theta_{init} (mm3 mm-3) is the moisture value at which the DSL initiates, \theta_{1} (mm3 mm-3) is the moisture value of the top model soil layer, and \theta_{air} (mm3 mm-3) is the ‘air dry’ soil moisture value (Dingman 2002):

(2.5.78)\theta_{air} = \Phi \left( \frac{\Psi_{sat}}{\Psi_{air}} \right)^{\frac{1}{B_{1}}} \ .

where \Phi is the porosity (mm3 mm-3), \Psi_{sat} is the saturated soil matric potential (mm), \Psi_{air} = 10^{7} mm is the air dry matric potential, and B_{1} is a function of soil texture (section 2.7.3.1).

The soil tortuosity is

(2.5.79)\tau = \Phi^{2}_{air}\left(\frac{\Phi_{air}}{\Phi}\right)^{\frac{3}{B_{1}}}

where \Phi_{air} (mm3 mm-3) is the air filled pore space

(2.5.80)\Phi_{air} = \Phi - \theta_{air} \ .

D_{v} depends on temperature

(2.5.81)D_{v} = 2.12 \times 10^{-5} \left(\frac{T_{1}}{T_{f}}\right)^{1.75} \ .

where T_{1} (K) is the temperature of the top soil layer and T_{f} (K) is the freezing temperature of water (Table 2.2.7).

The roughness lengths used to calculate r_{am} , r_{ah} , and r_{aw} are z_{0m} =z_{0m,\, g} , z_{0h} =z_{0h,\, g} , and z_{0w} =z_{0w,\, g} . The displacement height d=0. The momentum roughness length is z_{0m,\, g} =0.01 for soil, glaciers, and z_{0m,\, g} =0.0024 for snow-covered surfaces (f_{sno} >0). In general, z_{0m} is different from z_{0h} because the transfer of momentum is affected by pressure fluctuations in the turbulent waves behind the roughness elements, while for heat and water vapor transfer no such dynamical mechanism exists. Rather, heat and water vapor must be transferred by molecular diffusion across the interfacial sublayer. The following relation from Zilitinkevich (1970) is adopted by Zeng and Dickinson 1998

(2.5.82)z_{0h,\, g} =z_{0w,\, g} =z_{0m,\, g} e^{-a\left({u_{*} z_{0m,\, g} \mathord{\left/ {\vphantom {u_{*} z_{0m,\, g}  \upsilon }} \right. \kern-\nulldelimiterspace} \upsilon } \right)^{0.45} }

where the quantity {u_{\*} z_{0m,\, g} \mathord{\left/ {\vphantom {u_{*} z_{0m,\, g}  \upsilon }} \right. \kern-\nulldelimiterspace} \upsilon } is the roughness Reynolds number (and may be interpreted as the Reynolds number of the smallest turbulent eddy in the flow) with the kinematic viscosity of air \upsilon =1.5\times 10^{-5} m2 s-1 and a=0.13.

The numerical solution for the fluxes of momentum, sensible heat, and water vapor flux from non-vegetated surfaces proceeds as follows:

  1. An initial guess for the wind speed V_{a} is obtained from (2.5.24) assuming an initial convective velocity U_{c} =0 m s-1 for stable conditions (\theta _{v,\, atm} -\theta _{v,\, s} \ge 0 as evaluated from (2.5.50) ) and U_{c} =0.5 for unstable conditions (\theta _{v,\, atm} -\theta _{v,\, s} <0).

  2. An initial guess for the Monin-Obukhov length L is obtained from the bulk Richardson number using (2.5.46) and (2.5.48).

  3. The following system of equations is iterated three times:

  4. Friction velocity u_{*} ((2.5.32), (2.5.33), (2.5.34), (2.5.35))

  5. Potential temperature scale \theta _{*} ((2.5.37) , (2.5.38), (2.5.39), (2.5.40))

  6. Humidity scale q_{*} ((2.5.41), (2.5.42), (2.5.43), (2.5.44))

  7. Roughness lengths for sensible z_{0h,\, g} and latent heat z_{0w,\, g} ((2.5.82) )

  8. Virtual potential temperature scale \theta _{v*} ( (2.5.17))

  9. Wind speed including the convective velocity, V_{a} ( (2.5.24))

  10. Monin-Obukhov length L ((2.5.49))

  11. Aerodynamic resistances r_{am} , r_{ah} , and r_{aw} ((2.5.55), (2.5.56), (2.5.57))

  12. Momentum fluxes \tau _{x} , \tau _{y} ((2.5.5), (2.5.6))

  13. Sensible heat flux H_{g} ((2.5.62))

  14. Water vapor flux E_{g} ((2.5.66))

  15. 2-m height air temperature T_{2m} and specific humidity q_{2m} ((2.5.58) , (2.5.59))

The partial derivatives of the soil surface fluxes with respect to ground temperature, which are needed for the soil temperature calculations (section 2.6.1) and to update the soil surface fluxes (section 2.5.4), are

(2.5.83)\frac{\partial H_{g} }{\partial T_{g} } =\frac{\rho _{atm} C_{p} }{r_{ah} }

(2.5.84)\frac{\partial E_{g} }{\partial T_{g} } =\frac{\beta _{soi} \rho _{atm} }{r_{aw} } \frac{dq_{g} }{dT_{g} }

where

(2.5.85)\frac{dq_{g} }{dT_{g} } =\left(1-f_{sno} -f_{h2osfc} \right)\alpha _{soil} \frac{dq_{sat}^{T_{soil} } }{dT_{soil} } +f_{sno} \frac{dq_{sat}^{T_{sno} } }{dT_{sno} } +f_{h2osfc} \frac{dq_{sat}^{T_{h2osfc} } }{dT_{h2osfc} } .

The partial derivatives \frac{\partial r_{ah} }{\partial T_{g} } and \frac{\partial r_{aw} }{\partial T_{g} } , which cannot be determined analytically, are ignored for \frac{\partial H_{g} }{\partial T_{g} } and \frac{\partial E_{g} }{\partial T_{g} } .

2.5.3. Sensible and Latent Heat Fluxes and Temperature for Vegetated Surfaces

In the case of a vegetated surface, the sensible heat H and water vapor flux E are partitioned into vegetation and ground fluxes that depend on vegetation T_{v} and ground T_{g} temperatures in addition to surface temperature T_{s} and specific humidity q_{s} . Because of the coupling between vegetation temperature and fluxes, Newton-Raphson iteration is used to solve for the vegetation temperature and the sensible heat and water vapor fluxes from vegetation simultaneously using the ground temperature from the previous time step. In section 2.5.3.1, the equations used in the iteration scheme are derived. Details on the numerical scheme are provided in section 2.5.3.2.

2.5.3.1. Theory

The air within the canopy is assumed to have negligible capacity to store heat so that the sensible heat flux H between the surface at height z_{0h} +d and the atmosphere at height z_{atm,\, h} must be balanced by the sum of the sensible heat from the vegetation H_{v} and the ground H_{g}

(2.5.86)H=H_{v} +H_{g}

where, with reference to Figure 2.5.1,

(2.5.87)H=-\rho _{atm} C_{p} \frac{\left(\theta _{atm} -T_{s} \right)}{r_{ah} }

(2.5.88)H_{v} =-\rho _{atm} C_{p} \left(T_{s} -T_{v} \right)\frac{\left(L+S\right)}{r_{b} }

(2.5.89)H_{g} =\left(1-f_{sno} -f_{h2osfc} \right)H_{soil} +f_{sno} H_{snow} +f_{h2osfc} H_{h2osfc} \ ,

where

(2.5.90)H_{soil} =-\rho _{atm} C_{p} \frac{\left(T_{s} -T_{1} \right)}{r_{ah} ^{{'} } }

(2.5.91)H_{sno} =-\rho _{atm} C_{p} \frac{\left(T_{s} -T_{snl+1} \right)}{r_{ah} ^{{'} } }

(2.5.92)H_{h2osfc} =-\rho _{atm} C_{p} \frac{\left(T_{s} -T_{h2osfc} \right)}{r_{ah} ^{{'} } }

where \rho _{atm} is the density of atmospheric air (kg m-3), C_{p} is the specific heat capacity of air (J kg-1 K-1) (Table 2.2.7), \theta _{atm} is the atmospheric potential temperature (K), and r_{ah} is the aerodynamic resistance to sensible heat transfer (s m-1).

Here, T_{s} is the surface temperature at height z_{0h} +d, also referred to as the canopy air temperature. L and S are the exposed leaf and stem area indices (section 2.2.1.4), r_{b} is the leaf boundary layer resistance (s m-1), and r_{ah} ^{{'} } is the aerodynamic resistance (s m-1) to heat transfer between the ground at height z_{0h} ^{{'} } and the canopy air at height z_{0h} +d.

../../_images/image12.png

Figure 2.5.1 Figure Schematic diagram of sensible heat fluxes for (a) non-vegetated surfaces and (b) vegetated surfaces.

../../_images/image2.png

Figure 2.5.2 Figure Schematic diagram of water vapor fluxes for (a) non-vegetated surfaces and (b) vegetated surfaces.

Equations (2.5.86) - (2.5.89) can be solved for the canopy air temperature T_{s}

(2.5.93)T_{s} =\frac{c_{a}^{h} \theta _{atm} +c_{g}^{h} T_{g} +c_{v}^{h} T_{v} }{c_{a}^{h} +c_{g}^{h} +c_{v}^{h} }

where

(2.5.94)c_{a}^{h} =\frac{1}{r_{ah} }

(2.5.95)c_{g}^{h} =\frac{1}{r_{ah} ^{{'} } }

(2.5.96)c_{v}^{h} =\frac{\left(L+S\right)}{r_{b} }

are the sensible heat conductances from the canopy air to the atmosphere, the ground to canopy air, and leaf surface to canopy air, respectively (m s-1).

When the expression for T_{s} is substituted into equation (2.5.88), the sensible heat flux from vegetation H_{v} is a function of \theta _{atm} , T_{g} , and T_{v}

(2.5.97)H_{v} = -\rho _{atm} C_{p} \left[c_{a}^{h} \theta _{atm} +c_{g}^{h} T_{g} -\left(c_{a}^{h} +c_{g}^{h} \right)T_{v} \right]\frac{c_{v}^{h} }{c_{a}^{h} +c_{v}^{h} +c_{g}^{h} } .

Similarly, the expression for T_{s} can be substituted into equation to obtain the sensible heat flux from ground H_{g}

(2.5.98)H_{g} = -\rho _{atm} C_{p} \left[c_{a}^{h} \theta _{atm} +c_{v}^{h} T_{v} -\left(c_{a}^{h} +c_{v}^{h} \right)T_{g} \right]\frac{c_{g}^{h} }{c_{a}^{h} +c_{v}^{h} +c_{g}^{h} } .

The air within the canopy is assumed to have negligible capacity to store water vapor so that the water vapor flux E between the surface at height z_{0w} +d and the atmosphere at height z_{atm,\, w} must be balanced by the sum of the water vapor flux from the vegetation E_{v} and the ground E_{g}

(2.5.99)E = E_{v} +E_{g}

where, with reference to Figure 2.5.2,

(2.5.100)E = -\rho _{atm} \frac{\left(q_{atm} -q_{s} \right)}{r_{aw} }

(2.5.101)E_{v} = -\rho _{atm} \frac{\left(q_{s} -q_{sat}^{T_{v} } \right)}{r_{total} }

(2.5.102)E_{g} = \left(1-f_{sno} -f_{h2osfc} \right)E_{soil} +f_{sno} E_{snow} +f_{h2osfc} E_{h2osfc} \ ,

where

(2.5.103)E_{soil} = -\rho _{atm} \frac{\left(q_{s} -q_{soil} \right)}{r_{aw} ^{{'} } +r_{soil} }

(2.5.104)E_{sno} = -\rho _{atm} \frac{\left(q_{s} -q_{sno} \right)}{r_{aw} ^{{'} } +r_{soil} }

(2.5.105)E_{h2osfc} = -\rho _{atm} \frac{\left(q_{s} -q_{h2osfc} \right)}{r_{aw} ^{{'} } +r_{soil} }

where q_{atm} is the atmospheric specific humidity (kg kg-1), r_{aw} is the aerodynamic resistance to water vapor transfer (s m-1), q_{sat}^{T_{v} } (kg kg-1) is the saturation water vapor specific humidity at the vegetation temperature (section 2.5.5), q_{g} , q_{sno} , and q_{h2osfc} are the specific humidities of the soil, snow, and surface water (section 2.5.2), r_{aw} ^{{'} } is the aerodynamic resistance (s m-1) to water vapor transfer between the ground at height z_{0w} ^{{'} } and the canopy air at height z_{0w} +d, and r_{soil} ((2.5.76)) is a resistance to diffusion through the soil (s m-1). r_{total} is the total resistance to water vapor transfer from the canopy to the canopy air and includes contributions from leaf boundary layer and sunlit and shaded stomatal resistances r_{b} , r_{s}^{sun} , and r_{s}^{sha} (Figure 2.5.2). The water vapor flux from vegetation is the sum of water vapor flux from wetted leaf and stem area E_{v}^{w} (evaporation of water intercepted by the canopy) and transpiration from dry leaf surfaces E_{v}^{t}

(2.5.106)E_{v} =E_{v}^{w} +E_{v}^{t} .

Equations (2.5.99) - (2.5.102) can be solved for the canopy specific humidity q_{s}

(2.5.107)q_{s} =\frac{c_{a}^{w} q_{atm} +c_{g}^{w} q_{g} +c_{v}^{w} q_{sat}^{T_{v} } }{c_{a}^{w} +c_{v}^{w} +c_{g}^{w} }

where

(2.5.108)c_{a}^{w} =\frac{1}{r_{aw} }

(2.5.109)c_{v}^{w} =\frac{\left(L+S\right)}{r_{b} } r''

(2.5.110)c_{g}^{w} =\frac{1}{r_{aw} ^{{'} } +r_{soil} }

are the water vapor conductances from the canopy air to the atmosphere, the leaf to canopy air, and ground to canopy air, respectively. The term r'' is determined from contributions by wet leaves and transpiration and limited by available water and potential evaporation as

(2.5.111)r'' = \left\{
\begin{array}{lr}
\min \left(f_{wet} +r_{dry} ^{{'} {'} } ,\, \frac{E_{v}^{w,\, pot} r_{dry} ^{{'} {'} } +\frac{W_{can} }{\Delta t} }{E_{v}^{w,\, pot} } \right) & \qquad E_{v}^{w,\, pot} >0,\, \beta _{t} >0 \\
\min \left(f_{wet} ,\, \frac{E_{v}^{w,\, pot} r_{dry} ^{{'} {'} } +\frac{W_{can} }{\Delta t} }{E_{v}^{w,\, pot} } \right) & \qquad E_{v}^{w,\, pot} >0,\, \beta _{t} \le 0 \\
1 & \qquad E_{v}^{w,\, pot} \le 0
\end{array}\right\}

where f_{wet} is the fraction of leaves and stems that are wet (section 2.7.1), W_{can} is canopy water (kg m-2) (section 2.7.1), \Delta t is the time step (s), and \beta _{t} is a soil moisture function limiting transpiration (Chapter 2.9). The potential evaporation from wet foliage per unit wetted area is

(2.5.112)E_{v}^{w,\, pot} =-\frac{\rho _{atm} \left(q_{s} -q_{sat}^{T_{v} } \right)}{r_{b} } .

The term r_{dry} ^{{'} {'} } is

(2.5.113)r_{dry} ^{{'} {'} } =\frac{f_{dry} r_{b} }{L} \left(\frac{L^{sun} }{r_{b} +r_{s}^{sun} } +\frac{L^{sha} }{r_{b} +r_{s}^{sha} } \right)

where f_{dry} is the fraction of leaves that are dry (section 2.7.1), L^{sun} and L^{sha} are the sunlit and shaded leaf area indices (section 2.4.1), and r_{s}^{sun} and r_{s}^{sha} are the sunlit and shaded stomatal resistances (s m-1) (Chapter 2.9).

When the expression for q_{s} is substituted into equation (2.5.101), the water vapor flux from vegetation E_{v} is a function of q_{atm} , q_{g} , and q_{sat}^{T_{v} }

(2.5.114)E_{v} =-\rho _{atm} \left[c_{a}^{w} q_{atm} +c_{g}^{w} q_{g} -\left(c_{a}^{w} +c_{g}^{w} \right)q_{sat}^{T_{v} } \right]\frac{c_{v}^{w} }{c_{a}^{w} +c_{v}^{w} +c_{g}^{w} } .

Similarly, the expression for q_{s} can be substituted into (2.5.84) to obtain the water vapor flux from the ground beneath the canopy E_{g}

(2.5.115)E_{g} =-\rho _{atm} \left[c_{a}^{w} q_{atm} +c_{v}^{w} q_{sat}^{T_{v} } -\left(c_{a}^{w} +c_{v}^{w} \right)q_{g} \right]\frac{c_{g}^{w} }{c_{a}^{w} +c_{v}^{w} +c_{g}^{w} } .

The aerodynamic resistances to heat (moisture) transfer between the ground at height z_{0h} ^{{'} } (z_{0w} ^{{'} } ) and the canopy air at height z_{0h} +d (z_{0w} +d) are

(2.5.116)r_{ah} ^{{'} } =r_{aw} ^{{'} } =\frac{1}{C_{s} U_{av} }

where

(2.5.117)U_{av} =V_{a} \sqrt{\frac{1}{r_{am} V_{a} } } =u_{*}

is the magnitude of the wind velocity incident on the leaves (equivalent here to friction velocity) (m s-1) and C_{s} is the turbulent transfer coefficient between the underlying soil and the canopy air. C_{s} is obtained by interpolation between values for dense canopy and bare soil (Zeng et al. 2005)

(2.5.118)C_{s} =C_{s,\, bare} W+C_{s,\, dense} (1-W)

where the weight W is

(2.5.119)W=e^{-\left(L+S\right)} .

The dense canopy turbulent transfer coefficient (Dickinson et al. 1993) is

(2.5.120)C_{s,\, dense} =0.004 \ .

The bare soil turbulent transfer coefficient is

(2.5.121)C_{s,\, bare} =\frac{k}{a} \left(\frac{z_{0m,\, g} U_{av} }{\upsilon } \right)^{-0.45}

where the kinematic viscosity of air \upsilon =1.5\times 10^{-5} m2 s-1 and a=0.13.

The leaf boundary layer resistance r_{b} is

(2.5.122)r_{b} =\frac{1}{C_{v} } \left({U_{av} \mathord{\left/ {\vphantom {U_{av}  d_{leaf} }} \right. \kern-\nulldelimiterspace} d_{leaf} } \right)^{{-1\mathord{\left/ {\vphantom {-1 2}} \right. \kern-\nulldelimiterspace} 2} }

where C_{v} =0.01 ms-1/2 is the turbulent transfer coefficient between the canopy surface and canopy air, and d_{leaf} is the characteristic dimension of the leaves in the direction of wind flow (Table 2.5.1).

The partial derivatives of the fluxes from the soil beneath the canopy with respect to ground temperature, which are needed for the soil temperature calculations (section 2.6.1) and to update the soil surface fluxes (section 2.5.4), are

(2.5.123)\frac{\partial H_{g} }{\partial T_{g} } = \frac{\rho _{atm} C_{p} }{r'_{ah} } \frac{c_{a}^{h} +c_{v}^{h} }{c_{a}^{h} +c_{v}^{h} +c_{g}^{h} }

(2.5.124)\frac{\partial E_{g} }{\partial T_{g} } = \frac{\rho _{atm} }{r'_{aw} +r_{soil} } \frac{c_{a}^{w} +c_{v}^{w} }{c_{a}^{w} +c_{v}^{w} +c_{g}^{w} } \frac{dq_{g} }{dT_{g} } .

The partial derivatives \frac{\partial r'_{ah} }{\partial T_{g} } and \frac{\partial r'_{aw} }{\partial T_{g} } , which cannot be determined analytically, are ignored for \frac{\partial H_{g} }{\partial T_{g} } and \frac{\partial E_{g} }{\partial T_{g} } .

The roughness lengths used to calculate r_{am} , r_{ah} , and r_{aw} from (2.5.55), (2.5.56), and (2.5.57) are z_{0m} =z_{0m,\, v} , z_{0h} =z_{0h,\, v} , and z_{0w} =z_{0w,\, v} . The vegetation displacement height d and the roughness lengths are a function of plant height and adjusted for canopy density following Zeng and Wang (2007)

(2.5.125)z_{0m,\, v} = z_{0h,\, v} =z_{0w,\, v} =\exp \left[V\ln \left(z_{top} R_{z0m} \right)+\left(1-V\right)\ln \left(z_{0m,\, g} \right)\right]

(2.5.126)d = z_{top} R_{d} V

where z_{top} is canopy top height (m) (Table 2.2.2), R_{z0m} and R_{d} are the ratio of momentum roughness length and displacement height to canopy top height, respectively (Table 2.5.1), and z_{0m,\, g} is the ground momentum roughness length (m) (section 2.5.2). The fractional weight V is determined from

(2.5.127)V = \frac{1-\exp \left\{-\beta \min \left[L+S,\, \left(L+S\right)_{cr} \right]\right\}}{1-\exp \left[-\beta \left(L+S\right)_{cr} \right]}

where \beta =1 and \left(L+S\right)_{cr} = 2 (m2 m-2) is a critical value of exposed leaf plus stem area for which z_{0m} reaches its maximum.

Table 2.5.1 Plant functional type aerodynamic parameters

Plant functional type

R_{z0m}

R_{d}

d_{leaf} (m)

NET Temperate

0.055

0.67

0.04

NET Boreal

0.055

0.67

0.04

NDT Boreal

0.055

0.67

0.04

BET Tropical

0.075

0.67

0.04

BET temperate

0.075

0.67

0.04

BDT tropical

0.055

0.67

0.04

BDT temperate

0.055

0.67

0.04

BDT boreal

0.055

0.67

0.04

BES temperate

0.120

0.68

0.04

BDS temperate

0.120

0.68

0.04

BDS boreal

0.120

0.68

0.04

C3 arctic grass

0.120

0.68

0.04

C3 grass

0.120

0.68

0.04

C4 grass

0.120

0.68

0.04

Crop R

0.120

0.68

0.04

Crop I

0.120

0.68

0.04

Corn R

0.120

0.68

0.04

Corn I

0.120

0.68

0.04

Temp Cereal R

0.120

0.68

0.04

Temp Cereal I

0.120

0.68

0.04

Winter Cereal R

0.120

0.68

0.04

Winter Cereal I

0.120

0.68

0.04

Soybean R

0.120

0.68

0.04

Soybean I

0.120

0.68

0.04

2.5.3.2. Numerical Implementation

Canopy energy conservation gives

(2.5.128)-\overrightarrow{S}_{v} +\overrightarrow{L}_{v} \left(T_{v} \right)+H_{v} \left(T_{v} \right)+\lambda E_{v} \left(T_{v} \right)=0

where \overrightarrow{S}_{v} is the solar radiation absorbed by the vegetation (section 2.4.1), \overrightarrow{L}_{v} is the net longwave radiation absorbed by vegetation (section 2.4.2), and H_{v} and \lambda E_{v} are the sensible and latent heat fluxes from vegetation, respectively. The term \lambda is taken to be the latent heat of vaporization \lambda _{vap} (Table 2.2.7).

\overrightarrow{L}_{v} , H_{v} , and \lambda E_{v} depend on the vegetation temperature T_{v} . The Newton-Raphson method for finding roots of non-linear systems of equations can be applied to iteratively solve for T_{v} as

(2.5.129)\Delta T_{v} =\frac{\overrightarrow{S}_{v} -\overrightarrow{L}_{v} -H_{v} -\lambda E_{v} }{\frac{\partial \overrightarrow{L}_{v} }{\partial T_{v} } +\frac{\partial H_{v} }{\partial T_{v} } +\frac{\partial \lambda E_{v} }{\partial T_{v} } }

where \Delta T_{v} =T_{v}^{n+1} -T_{v}^{n} and the subscript “n” indicates the iteration.

The partial derivatives are

(2.5.130)\frac{\partial \overrightarrow{L}_{v} }{\partial T_{v} } =4\varepsilon _{v} \sigma \left[2-\varepsilon _{v} \left(1-\varepsilon _{g} \right)\right]T_{v}^{3}

(2.5.131)\frac{\partial H_{v} }{\partial T_{v} } =\rho _{atm} C_{p} \left(c_{a}^{h} +c_{g}^{h} \right)\frac{c_{v}^{h} }{c_{a}^{h} +c_{v}^{h} +c_{g}^{h} }

(2.5.132)\frac{\partial \lambda E_{v} }{\partial T_{v} } =\lambda \rho _{atm} \left(c_{a}^{w} +c_{g}^{w} \right)\frac{c_{v}^{w} }{c_{a}^{w} +c_{v}^{w} +c_{g}^{w} } \frac{dq_{sat}^{T_{v} } }{dT_{v} } .

The partial derivatives \frac{\partial r_{ah} }{\partial T_{v} } and \frac{\partial r_{aw} }{\partial T_{v} } , which cannot be determined analytically, are ignored for \frac{\partial H_{v} }{\partial T_{v} } and \frac{\partial \lambda E_{v} }{\partial T_{v} } . However, if \zeta changes sign more than four times during the temperature iteration, \zeta =-0.01. This helps prevent “flip-flopping” between stable and unstable conditions. The total water vapor flux E_{v} , transpiration flux E_{v}^{t} , and sensible heat flux H_{v} are updated for changes in leaf temperature as

(2.5.133)E_{v} =-\rho _{atm} \left[c_{a}^{w} q_{atm} +c_{g}^{w} q_{g} -\left(c_{a}^{w} +c_{g}^{w} \right)\left(q_{sat}^{T_{v} } +\frac{dq_{sat}^{T_{v} } }{dT_{v} } \Delta T_{v} \right)\right]\frac{c_{v}^{w} }{c_{a}^{w} +c_{v}^{w} +c_{g}^{w} }

(2.5.134)E_{v}^{t} =-r_{dry} ^{{'} {'} } \rho _{atm} \left[c_{a}^{w} q_{atm} +c_{g}^{w} q_{g} -\left(c_{a}^{w} +c_{g}^{w} \right)\left(q_{sat}^{T_{v} } +\frac{dq_{sat}^{T_{v} } }{dT_{v} } \Delta T_{v} \right)\right]\frac{c_{v}^{h} }{c_{a}^{w} +c_{v}^{w} +c_{g}^{w} }

(2.5.135)H_{v} =-\rho _{atm} C_{p} \left[c_{a}^{h} \theta _{atm} +c_{g}^{h} T_{g} -\left(c_{a}^{h} +c_{g}^{h} \right)\left(T_{v} +\Delta T_{v} \right)\right]\frac{c_{v}^{h} }{c_{a}^{h} +c_{v}^{h} +c_{g}^{h} } .

The numerical solution for vegetation temperature and the fluxes of momentum, sensible heat, and water vapor flux from vegetated surfaces proceeds as follows:

  1. Initial values for canopy air temperature and specific humidity are obtained from

    (2.5.136)T_{s} =\frac{T_{g} +\theta _{atm} }{2}

    (2.5.137)q_{s} =\frac{q_{g} +q_{atm} }{2} .

  2. An initial guess for the wind speed V_{a} is obtained from (2.5.24) assuming an initial convective velocity U_{c} =0 m s-1 for stable conditions (\theta _{v,\, atm} -\theta _{v,\, s} \ge 0 as evaluated from (2.5.50) ) and U_{c} =0.5 for unstable conditions (\theta _{v,\, atm} -\theta _{v,\, s} <0).

  3. An initial guess for the Monin-Obukhov length L is obtained from the bulk Richardson number using equation and (2.5.46) and (2.5.48).

  4. Iteration proceeds on the following system of equations:

  5. Friction velocity u_{*} ((2.5.32), (2.5.33), (2.5.34), (2.5.35))

  6. Ratio \frac{\theta _{*} }{\theta _{atm} -\theta _{s} } ((2.5.37) , (2.5.38), (2.5.39), (2.5.40))

  7. Ratio \frac{q_{*} }{q_{atm} -q_{s} } ((2.5.41), (2.5.42), (2.5.43), (2.5.44))

  8. Aerodynamic resistances r_{am} , r_{ah} , and r_{aw} ((2.5.55), (2.5.56), (2.5.57))

  9. Magnitude of the wind velocity incident on the leaves U_{av} ((2.5.117) )

  10. Leaf boundary layer resistance r_{b} ((2.5.136) )

  11. Aerodynamic resistances r_{ah} ^{{'} } and r_{aw} ^{{'} } ((2.5.116) )

  12. Sunlit and shaded stomatal resistances r_{s}^{sun} and r_{s}^{sha} (Chapter 2.9)

  13. Sensible heat conductances c_{a}^{h} , c_{g}^{h} , and c_{v}^{h} ((2.5.94), (2.5.95), (2.5.96))

  14. Latent heat conductances c_{a}^{w} , c_{v}^{w} , and c_{g}^{w} ((2.5.108), (2.5.109), (2.5.110))

  15. Sensible heat flux from vegetation H_{v} ((2.5.97) )

  16. Latent heat flux from vegetation \lambda E_{v} ((2.5.101) )

  17. If the latent heat flux has changed sign from the latent heat flux computed at the previous iteration (\lambda E_{v} ^{n+1} \times \lambda E_{v} ^{n} <0), the latent heat flux is constrained to be 10% of the computed value. The difference between the constrained and computed value (\Delta _{1} =0.1\lambda E_{v} ^{n+1} -\lambda E_{v} ^{n+1} ) is added to the sensible heat flux later.

  18. Change in vegetation temperature \Delta T_{v} ((2.5.129) ) and update the vegetation temperature as T_{v}^{n+1} =T_{v}^{n} +\Delta T_{v} . T_{v} is constrained to change by no more than 1ºK in one iteration. If this limit is exceeded, the energy error is

    (2.5.138)\Delta _{2} =\overrightarrow{S}_{v} -\overrightarrow{L}_{v} -\frac{\partial \overrightarrow{L}_{v} }{\partial T_{v} } \Delta T_{v} -H_{v} -\frac{\partial H_{v} }{\partial T_{v} } \Delta T_{v} -\lambda E_{v} -\frac{\partial \lambda E_{v} }{\partial T_{v} } \Delta T_{v}

where \Delta T_{v} =1{\rm \; or\; }-1. The error \Delta _{2} is added to the sensible heat flux later.

  1. Water vapor flux E_{v} ((2.5.133) )

  2. Transpiration E_{v}^{t} ((2.5.134) if \beta_{t} >0, otherwise E_{v}^{t} =0)

  3. The water vapor flux E_{v} is constrained to be less than or equal to the sum of transpiration E_{v}^{t} and the water available from wetted leaves and stems {W_{can} \mathord{\left/ {\vphantom {W_{can}  \Delta t}} \right. \kern-\nulldelimiterspace} \Delta t} . The energy error due to this constraint is

    (2.5.139)\Delta _{3} =\max \left(0,\, E_{v} -E_{v}^{t} -\frac{W_{can} }{\Delta t} \right).

The error \lambda \Delta _{3} is added to the sensible heat flux later.

  1. Sensible heat flux H_{v} ((2.5.135) ). The three energy error terms, \Delta _{1} , \Delta _{2} , and \lambda \Delta _{3} are also added to the sensible heat flux.

  2. The saturated vapor pressure e_{i} (Chapter 2.9), saturated specific humidity q_{sat}^{T_{v} } and its derivative \frac{dq_{sat}^{T_{v} } }{dT_{v} } at the leaf surface (section 2.5.5), are re-evaluated based on the new T_{v} .

  3. Canopy air temperature T_{s} ((2.5.93) )

  4. Canopy air specific humidity q_{s} ((2.5.107) )

  5. Temperature difference \theta _{atm} -\theta _{s}

  6. Specific humidity difference q_{atm} -q_{s}

  7. Potential temperature scale \theta _{*} =\frac{\theta _{*} }{\theta _{atm} -\theta _{s} } \left(\theta _{atm} -\theta _{s} \right) where \frac{\theta _{*} }{\theta _{atm} -\theta _{s} } was calculated earlier in the iteration

  8. Humidity scale q_{*} =\frac{q_{*} }{q_{atm} -q_{s} } \left(q_{atm} -q_{s} \right) where \frac{q_{*} }{q_{atm} -q_{s} } was calculated earlier in the iteration

  9. Virtual potential temperature scale \theta _{v*} ((2.5.17) )

  10. Wind speed including the convective velocity, V_{a} ((2.5.24) )

  11. Monin-Obukhov length L ((2.5.49) )

  12. The iteration is stopped after two or more steps if \tilde{\Delta }T_{v} <0.01 and \left|\lambda E_{v}^{n+1} -\lambda E_{v}^{n} \right|<0.1 where \tilde{\Delta }T_{v} =\max \left(\left|T_{v}^{n+1} -T_{v}^{n} \right|,\, \left|T_{v}^{n} -T_{v}^{n-1} \right|\right), or after forty iterations have been carried out.

  13. Momentum fluxes \tau _{x} , \tau _{y} ((2.5.5), (2.5.6))

  14. Sensible heat flux from ground H_{g} ((2.5.89) )

  15. Water vapor flux from ground E_{g} ((2.5.102) )

  16. 2-m height air temperature T_{2m} , specific humidity q_{2m} , relative humidity RH_{2m} ((2.5.58) , (2.5.59), (2.5.60))

2.5.4. Update of Ground Sensible and Latent Heat Fluxes

The sensible and water vapor heat fluxes derived above for bare soil and soil beneath canopy are based on the ground surface temperature from the previous time step T_{g}^{n} and are used as the surface forcing for the solution of the soil temperature equations (section 2.6.1). This solution yields a new ground surface temperature T_{g}^{n+1} . The ground sensible and water vapor fluxes are then updated for T_{g}^{n+1} as

(2.5.140)H'_{g} =H_{g} +\left(T_{g}^{n+1} -T_{g}^{n} \right)\frac{\partial H_{g} }{\partial T_{g} }

(2.5.141)E'_{g} =E_{g} +\left(T_{g}^{n+1} -T_{g}^{n} \right)\frac{\partial E_{g} }{\partial T_{g} }

where H_{g} and E_{g} are the sensible heat and water vapor fluxes derived from equations and for non-vegetated surfaces and equations and for vegetated surfaces using T_{g}^{n} . One further adjustment is made to H'_{g} and E'_{g} . If the soil moisture in the top snow/soil layer is not sufficient to support the updated ground evaporation, i.e., if E'_{g} > 0 and f_{evap} < 1 where

(2.5.142)f_{evap} =\frac{{\left(w_{ice,\; snl+1} +w_{liq,\, snl+1} \right)\mathord{\left/ {\vphantom {\left(w_{ice,\; snl+1} +w_{liq,\, snl+1} \right) \Delta t}} \right. \kern-\nulldelimiterspace} \Delta t} }{\sum _{j=1}^{npft}\left(E'_{g} \right)_{j} \left(wt\right)_{j}  } \le 1,

an adjustment is made to reduce the ground evaporation accordingly as

(2.5.143)E''_{g} =f_{evap} E'_{g} .

The term \sum _{j=1}^{npft}\left(E'_{g} \right)_{j} \left(wt\right)_{j} is the sum of E'_{g} over all evaporating PFTs where \left(E'_{g} \right)_{j} is the ground evaporation from the j^{th} PFT on the column, \left(wt\right)_{j} is the relative area of the j^{th} PFT with respect to the column, and npft is the number of PFTs on the column. w_{ice,\, snl+1} and w_{liq,\, snl+1} are the ice and liquid water contents (kg m-2) of the top snow/soil layer (Chapter 2.7). Any resulting energy deficit is assigned to sensible heat as

(2.5.144)H''_{g} =H_{g} +\lambda \left(E'_{g} -E''_{g} \right).

The ground water vapor flux E''_{g} is partitioned into evaporation of liquid water from snow/soil q_{seva} (kgm-2 s-1), sublimation from snow/soil ice q_{subl} (kg m-2 s-1), liquid dew on snow/soil q_{sdew} (kg m-2 s-1), or frost on snow/soil q_{frost} (kg m-2 s-1) as

(2.5.145)q_{seva} =\max \left(E''_{sno} \frac{w_{liq,\, snl+1} }{w_{ice,\; snl+1} +w_{liq,\, snl+1} } ,0\right)\qquad E''_{sno} \ge 0,\, w_{ice,\; snl+1} +w_{liq,\, snl+1} >0

(2.5.146)q_{subl} =E''_{sno} -q_{seva} \qquad E''_{sno} \ge 0

(2.5.147)q_{sdew} =\left|E''_{sno} \right|\qquad E''_{sno} <0{\rm \; and\; }T_{g} \ge T_{f}

(2.5.148)q_{frost} =\left|E''_{sno} \right|\qquad E''_{sno} <0{\rm \; and\; }T_{g} <T_{f} .

The loss or gain in snow mass due to q_{seva} , q_{subl} , q_{sdew} , and q_{frost} on a snow surface are accounted for during the snow hydrology calculations (Chapter 2.8). The loss of soil and surface water due to q_{seva} is accounted for in the calculation of infiltration (section 2.7.2.3), while losses or gains due to q_{subl} , q_{sdew} , and q_{frost} on a soil surface are accounted for following the sub-surface drainage calculations (section 2.7.5).

The ground heat flux G is calculated as

(2.5.149)G=\overrightarrow{S}_{g} -\overrightarrow{L}_{g} -H_{g} -\lambda E_{g}

where \overrightarrow{S}_{g} is the solar radiation absorbed by the ground (section 2.4.1), \overrightarrow{L}_{g} is the net longwave radiation absorbed by the ground (section 2.4.2)

(2.5.150)\vec{L}_{g} =L_{g} \uparrow -\delta _{veg} \varepsilon _{g} L_{v} \, \downarrow -\left(1-\delta _{veg} \right)\varepsilon _{g} L_{atm} \, \downarrow +4\varepsilon _{g} \sigma \left(T_{g}^{n} \right)^{3} \left(T_{g}^{n+1} -T_{g}^{n} \right),

where

(2.5.151)L_{g} \uparrow =\varepsilon _{g} \sigma \left[\left(1-f_{sno} -f_{h2osfc} \right)\left(T_{1}^{n} \right)^{4} +f_{sno} \left(T_{sno}^{n} \right)^{4} +f_{h2osfc} \left(T_{h2osfc}^{n} \right)^{4} \right]

and H_{g} and \lambda E_{g} are the sensible and latent heat fluxes after the adjustments described above.

When converting ground water vapor flux to an energy flux, the term \lambda is arbitrarily assumed to be

(2.5.152)\lambda =\left\{\begin{array}{l} {\lambda _{sub} \qquad {\rm if\; }w_{liq,\, snl+1} =0{\rm \; and\; }w_{ice,\, snl+1} >0} \\ {\lambda _{vap} \qquad {\rm otherwise}} \end{array}\right\}

where \lambda _{sub} and \lambda _{vap} are the latent heat of sublimation and vaporization, respectively (J (kg-1) (Table 2.2.7). When converting vegetation water vapor flux to an energy flux, \lambda _{vap} is used.

The system balances energy as

(2.5.153)\overrightarrow{S}_{g} +\overrightarrow{S}_{v} +L_{atm} \, \downarrow -L\, \uparrow -H_{v} -H_{g} -\lambda _{vap} E_{v} -\lambda E_{g} -G=0.

2.5.5. Saturation Vapor Pressure

Saturation vapor pressure e_{sat}^{T} (Pa) and its derivative \frac{de_{sat}^{T} }{dT} , as a function of temperature T (ºC), are calculated from the eighth-order polynomial fits of Flatau et al. (1992)

(2.5.154)e_{sat}^{T} =100\left[a_{0} +a_{1} T+\cdots +a_{n} T^{n} \right]

(2.5.155)\frac{de_{sat}^{T} }{dT} =100\left[b_{0} +b_{1} T+\cdots +b_{n} T^{n} \right]

where the coefficients for ice are valid for -75\, ^{\circ } {\rm C}\le T<0\, ^{\circ } {\rm C} and the coefficients for water are valid for 0\, ^{\circ } {\rm C}\le T\le 100\, ^{\circ } {\rm C} (Table 2.5.2 and Table 2.5.3). The saturated water vapor specific humidity q_{sat}^{T} and its derivative \frac{dq_{sat}^{T} }{dT} are

(2.5.156)q_{sat}^{T} =\frac{0.622e_{sat}^{T} }{P_{atm} -0.378e_{sat}^{T} }

(2.5.157)\frac{dq_{sat}^{T} }{dT} =\frac{0.622P_{atm} }{\left(P_{atm} -0.378e_{sat}^{T} \right)^{2} } \frac{de_{sat}^{T} }{dT} .

Table 2.5.2 Coefficients for e_{sat}^{T}

water

ice

a_{0}

6.11213476

6.11123516

a_{1}

4.44007856 \times 10^{-1} | 5.03109514\times 10^{-1}

a_{2}

1.43064234 \times 10^{-2} | 1.88369801\times 10^{-2}

a_{3}

2.64461437 \times 10^{-4} | 4.20547422\times 10^{-4}

a_{4}

3.05903558 \times 10^{-6} | 6.14396778\times 10^{-6}

a_{5}

1.96237241 \times 10^{-8} | 6.02780717\times 10^{-8}

a_{6}

8.92344772 \times 10^{-11} | 3.87940929\times 10^{-10}

a_{7}

-3.73208410 \times 10^{-13} | 1.49436277\times 10^{-12}

a_{8}

2.09339997 \times 10^{-16} | 2.62655803\times 10^{-15}

Table 2.5.3 Coefficients for \frac{de_{sat}^{T} }{dT}

water

ice

b_{0}

4.44017302\times 10^{-1}

5.03277922\times 10^{-1}

b_{1}

2.86064092\times 10^{-2}

3.77289173\times 10^{-2}

b_{2}

7.94683137\times 10^{-4}

1.26801703\times 10^{-3}

b_{3}

1.21211669\times 10^{-5}

2.49468427\times 10^{-5}

b_{4}

1.03354611\times 10^{-7}

3.13703411\times 10^{-7}

b_{5}

4.04125005\times 10^{-10}

2.57180651\times 10^{-9}

b_{6}

-7.88037859 \times 10^{-13}

1.33268878\times 10^{-11}

b_{7}

-1.14596802 \times 10^{-14}

3.94116744\times 10^{-14}

b_{8}

3.81294516\times 10^{-17}

4.98070196\times 10^{-17}