To install click the Add extension button. That's it.

The source code for the WIKI 2 extension is being checked by specialists of the Mozilla Foundation, Google, and Apple. You could also do it yourself at any point in time.

4,5
Kelly Slayton
Congratulations on this excellent venture… what a great idea!
Alexander Grigorievskiy
I use WIKI 2 every day and almost forgot how the original Wikipedia looks like.
Live Statistics
English Articles
Improved in 24 Hours
Added in 24 Hours
Languages
Recent
Show all languages
What we do. Every page goes through several hundred of perfecting techniques; in live mode. Quite the same Wikipedia. Just better.
.
Leo
Newton
Brights
Milds

Penman–Monteith equation

From Wikipedia, the free encyclopedia

The Penman–Monteith equation approximates net evapotranspiration (ET) from meteorological data, as a replacement for direct measurement of evapotranspiration. The equation is widely used, and was derived by the United Nations Food and Agriculture Organization for modeling reference evapotranspiration ET0.[1]

Significance

Evapotranspiration contributions are very significant in a watershed's water balance, yet are often not emphasized in results because the precision of this component is often weak relative to more directly measured phenomena, e.g. rain and stream flow. In addition to weather uncertainties, the Penman–Monteith equation is sensitive to vegetation specific parameters, e.g. stomatal resistance or conductance.[2]

Various forms of crop coefficients (Kc) account for differences between specific vegetation modeled and a reference evapotranspiration (RET or ET0) standard. Stress coefficients (Ks) account for reductions in ET due to environmental stress (e.g. soil saturation reduces root-zone O2, low soil moisture induces wilt, air pollution effects, and salinity). Models of native vegetation cannot assume crop management to avoid recurring stress.

Equation

Per Monteith's Evaporation and Environment,[3] the equation is:

λv = Latent heat of vaporization. Energy required per unit mass of water vaporized. (J g−1)
Lv = Volumetric latent heat of vaporization. Energy required per water volume vaporized. (Lv = 2453 MJ m−3)
E = Mass water evapotranspiration rate (g s−1 m−2)
ET = Water volume evapotranspired (mm s−1)
Δ = Rate of change of saturation specific humidity with air temperature. (Pa K−1)
Rn = Net irradiance (W m−2), the external source of energy flux
G = Ground heat flux (W m−2), usually difficult to measure
cp = Specific heat capacity of air (J kg−1 K−1)
ρa = dry air density (kg m−3)
δe = vapor pressure deficit (Pa)
ga = Conductivity of air, atmospheric conductance (m s−1)
gs = Conductivity of stoma, surface or stomatal conductance (m s−1)
γ = Psychrometric constant (γ ≈ 66 Pa K−1)

Note: Often resistances are used rather than conductivities.

where rc refers to the resistance to flux from a vegetation canopy to the extent of some defined boundary layer.

The atmospheric conductance ga accounts for aerodynamic effects like the zero plane displacement height and the roughness length of the surface. The stomatal conductance gs accounts for effect of leaf density (Leaf Area Index), water stress and CO2 concentration in the air, that is to say to plant reaction to external factors. Different models exist to link the stomatal conductance to these vegetation characteristics, like the ones from P.G. Jarvis (1976) [4] or Jacobs et al. (1996).[5]

Accuracy

While the Penman-Monteith method is widely considered accurate for practical purposes and is recommended by the Food and Agriculture Organization of the United Nations,[1] errors when compared to direct measurement or other techniques can range from -9 to 40%.[6]

Variations and alternatives

FAO 56 Penman-Monteith equation

To avoid the inherent complexity of determining stomatal and atmospheric conductance, the Food and Agriculture Organization proposed in 1998 [1] a simplified equation for the reference evapotranspiration ET0. It is defined as the evapotranpiration for "[an] hypothetical reference crop with an assumed crop height of 0.12 m, a fixed surface resistance of 70 s m-1 and an albedo of 0.23." This reference surface is defined to represent "an extensive surface of green grass of uniform height, actively growing, completely shading the ground and with adequate water". The corresponding equation is:

ET0 = Reference evapotranspiration, Water volume evapotranspired (mm day−1)
Δ = Rate of change of saturation specific humidity with air temperature. (Pa K−1)
Rn = Net irradiance (MJ m−2 day−1), the external source of energy flux
G = Ground heat flux (MJ m−2 day−1), usually equivalent to zero on a day
T = Air temperature at 2m (K)
u2 = Wind speed at 2m height (m/s)
δe = vapor pressure deficit (kPa)
γ = Psychrometric constant (γ ≈ 66 Pa K−1)

N.B.: The coefficient 0.408 and 900 are not unitless but account for the conversion from energy values to equivalent water depths: radiation [mm day−1] = 0.408 radiation [MJ m−2 day−1].

This reference evapotranspiration ET0 can then be used to evaluate the evapotranspiration rate ET from unstressed plant through crop coefficients Kc: ET = Kc * ET0.[1]

Variations

The standard methods of the American Society of Civil Engineers modify the standard Penman–Monteith equation for use with an hourly time step. The SWAT model is one of many GIS-integrated hydrologic models estimating ET using Penman–Monteith equations.[7]

Priestley–Taylor

The Priestley–Taylor equation was developed as a substitute to the Penman–Monteith equation to remove dependence on observations. For Priestley–Taylor, only radiation (irradiance) observations are required. This is done by removing the aerodynamic terms from the Penman–Monteith equation and adding an empirically derived constant factor, .

The underlying concept behind the Priestley–Taylor model is that an air mass moving above a vegetated area with abundant water would become saturated with water. In these conditions, the actual evapotranspiration would match the Penman rate of reference evapotranspiration. However, observations revealed that actual evaporation was 1.26 times greater than reference evaporation, and therefore the equation for actual evaporation was found by taking reference evapotranspiration and multiplying it by .[8] The assumption here is for vegetation with an abundant water supply (i.e. the plants have low moisture stress). Areas like arid regions with high moisture stress are estimated to have higher values.[9]

The assumption that an air mass moving over a vegetated surface with abundant water saturates has been questioned later. The lowest and turbulent part of the atmosphere, the atmospheric boundary layer, is not a closed box, but constantly brings in dry air from higher up in the atmosphere towards the surface. As water evaporates more easily into a dry atmosphere, evapotranspiration is enhanced. This explains the larger than unity value of the Priestley-Taylor parameter . The proper equilibrium of the system has been derived and involves the characteristics of the interface of the atmospheric boundary layer and the overlying free atmosphere.[10][11]

History

The equation is named after Howard Penman and John Monteith. Penman first published his equation in 1948 and Monteith revised it in 1965.[3]

References

  1. ^ a b c d Richard G. Allen; Luis S. Pereira; Dirk Raes; Martin Smith (1998). Crop Evapotranspiration – Guidelines for Computing Crop Water Requirements. FAO Irrigation and drainage paper 56. Rome, Italy: Food and Agriculture Organization of the United Nations. ISBN 978-92-5-104219-9.
  2. ^ Keith Beven (1979). "A sensitivity analysis of the Penman–Monteith actual evapotranspiration estimates". Journal of Hydrology. 44 (3–4): 169–190. Bibcode:1979JHyd...44..169B. doi:10.1016/0022-1694(79)90130-6.
  3. ^ a b Monteith, J. L. (1965). "Evaporation and environment". Symposia of the Society for Experimental Biology. 19: 205–234. PMID 5321565.
  4. ^ Jarvis, P. (1976). "The interpretation of the variations in leaf water potential and stomatal conductance found in canopies in the field". Philosophical Transactions of the Royal Society of London. B, Biological Sciences. 273 (927): 593–610. doi:10.1098/rstb.1976.0035.
  5. ^ Jacobs, C.M.J (1996). "Stomatal behaviour and photosynthetic rate of unstressed grapevines in semi-arid conditions". Agricultural and Forest Meteorology. 80 (2–4): 111–134. doi:10.1016/0168-1923(95)02295-3.
  6. ^ Widmoser, Peter (2009-04-01). "A discussion on and alternative to the Penman–Monteith equation". Agricultural Water Management. 96 (4): 711–721. doi:10.1016/j.agwat.2008.10.003. ISSN 0378-3774.
  7. ^ "Hydrology Models in GRASS". 2007-07-03. Archived from the original on 3 July 2007. Retrieved 2022-02-21.
  8. ^ Priestley, C. H. B.; Taylor, R. J. (1972-02-01). "On the Assessment of Surface Heat Flux and Evaporation Using Large-Scale Parameters". Monthly Weather Review. 100 (2): 81–92. doi:10.1175/1520-0493(1972)100<0081:OTAOSH>2.3.CO;2. ISSN 1520-0493.
  9. ^ M. E. Jensen, R. D. Burman & R. G. Allen, ed. (1990). Evapotranspiration and Irrigation Water Requirement. ASCE Manuals and Reports on Engineering Practices. Vol. 70. New York, NY: American Society of Civil Engineers. ISBN 978-0-87262-763-5.
  10. ^ Culf, A. (1994). "Equilibrium evaporation beneath a growing convective boundary layer". Boundary-Layer Meteorology. 70 (1–2): 34–49. Bibcode:1994BoLMe..70...37C. doi:10.1007/BF00712522. S2CID 123108265.
  11. ^ van Heerwaarden, C. C.; et al. (2009). "Interactions between dry-air entrainment, surface evaporation and convective boundary layer development". Quarterly Journal of the Royal Meteorological Society. 135 (642): 1277–1291. Bibcode:2009QJRMS.135.1277V. doi:10.1002/qj.431. S2CID 123228410.

External links

This page was last edited on 12 February 2024, at 19:20
Basis of this page is in Wikipedia. Text is available under the CC BY-SA 3.0 Unported License. Non-text media are available under their specified licenses. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc. WIKI 2 is an independent company and has no affiliation with Wikimedia Foundation.