Skip to main content
Log in

Semi-analytical Footprint Model Compliant with Arbitrary Atmospheric Stratification: Application to Monin–Obukhov Profiles

  • Research Article
  • Published:
Boundary-Layer Meteorology Aims and scope Submit manuscript

Abstract

A new model is proposed for the so-called scalar footprint and flux footprint in the atmospheric boundary layer. The underlying semi-analytical model allows computing the scalar concentration and flux fields related to turbulent diffusion of heat, water-vapor or to the dispersion of any scalar (e.g. passive pollutant) in the framework of K-theory. It offers improved capabilities regarding the representation of the gradual stratification in the boundary layer. In this model, the boundary layer is split in a series of sublayers in which the aerodynamic inertivity (a compound parameter aggregating wind-speed and eddy-diffusivity) is approximated by a sum of two power-law functions of a new vertical scale corresponding to the height-dependent downwind extension of the plume. This multilayer approach allows fitting with vanishing error any boundary-layer stratification, in particular those described by the Monin–Obukhov similarity theory (MOST) in the surface layer, while keeping the computation time of the footprint to low values. As a complement, a fully analytical surrogate model is presented for practical applications. For MOST profiles, the flux (resp. concentration) footprint is, to a RMS difference less than 1% (resp. 1.2%), equal (resp. equal to a constant multiplicative factor) to the inverse Gamma distribution. The optimal parameters of this distribution were evaluated for a broad range of atmospheric conditions and height. Regression formulas were also provided to compute the crosswind-integrated flux footprint distribution easily and with less than 1.6% RMS residual error. A comparison with the well-known footprint approximate model by Kormann and Meixner and the one by Hsieh, Katul and Chi has allowed quantifying their performances and limitations.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15

Similar content being viewed by others

References

  • Bai J, Jia L, Liu S, Xu Z, Hu G, Zhu M, Song L (2015) Characterizing the footprint of eddy covariance system and large aperture scintillometer measurements to validate satellite-based surface fluxes. IEEE Geosci Remote Sens Lett 12(5):943–947

  • Baldocchi D, Falge E, Gu L, Olson R, Hollinger D, Running S, Anthoni P, Bernhofer C, Davis K, Evans R et al (2001) Fluxnet: a new tool to study the temporal and spatial variability of ecosystem-scale carbon dioxide, water vapor, and energy flux densities. Bull Am Meteorol Soc 82(11):2415–2434

    Article  Google Scholar 

  • Basu S, Lacser A (2017) A cautionary note on the use of Monin–Obukhov similarity theory in very high-resolution large-eddy simulations. Bound Layer Meteorol 163(2):351–355

    Article  Google Scholar 

  • Businger JA, Wyngaard JC, Izumi Y, Bradley EF (1971) Flux-profile relationships in the atmospheric surface layer. J Atmos Sci 28(2):181–189

    Article  Google Scholar 

  • Buske D, Vilhena MT, Moreira DM, Tirabassi T (2007) An analytical solution of the advection–diffusion equation considering non-local turbulence closure. Environ Fluid Mech 7(1):43–54

    Article  Google Scholar 

  • Buske D, Vilhena MT, Moreira DM, Tirabassi T (2007) Simulation of pollutant dispersion for low wind conditions in stable and convective planetary boundary layer. Atmos Environ 41(26):5496–5501

    Article  Google Scholar 

  • Buske D, Vilhena MT, Bodmann B, Tirabassi T (2012) Analytical model for air pollution in the atmospheric boundary layer. In: Khare M (ed) Air pollution: monitoring, modelling and health. IntechOpen, pp 17–34

  • Buske D, Vilhena MT, Tirabassi T, Bodmann B (2012b) Air pollution steady-state advection–diffusion equation: the general three-dimensional solution. J Environ Prot 2012

  • Calder K (1949) Eddy diffusion and evaporation in flow over aerodynamically smooth and rough surfaces: a treatment based on laboratory laws of turbulent flow with special reference to conditions in the lower atmosphere. Q J Mech Appl Math 2(2):153–176

    Article  Google Scholar 

  • Carvalho JC, Vilhena MT, Moreira DM (2007) Comparison between Eulerian and Lagrangian semi-analytical models to simulate the pollutant dispersion in the PBL. Appl Math Model 31(1):120–129

    Article  Google Scholar 

  • Chu H, Luo X, Ouyang Z, Chan WS, Dengel S, Biraud SC, Torn MS, Metzger S, Kumar J, Arain MA et al (2021) Representativeness of Eddy-Covariance flux footprints for areas surrounding AmeriFlux sites. Agric For Meteorol 301(108):350

    Google Scholar 

  • Costa CP, Vilhena MT, Moreira DM, Tirabassi T (2006) Semi-analytical solution of the steady three-dimensional advection–diffusion equation in the planetary boundary layer. Atmos Environ 40(29):5659–5669

    Article  Google Scholar 

  • Costa CP, Rui K, Pérez-Fernández LD (2018) Different numerical inversion algorithms of the Laplace transform for the solution of the advection–diffusion equation with non-local closure in air pollution modeling. TEMA (São Carlos) 19(1):43–58

    Article  Google Scholar 

  • De Hoog FR, Knight J, Stokes A (1982) An improved method for numerical inversion of Laplace transforms. SIAM J Sci Stat Comput 3(3):357–366

    Article  Google Scholar 

  • De Ridder K (2010) Bulk transfer relations for the roughness sublayer. Bound Layer Meteorol 134(2):257–267

    Article  Google Scholar 

  • Degrazia GA, Moreira DM, Vilhena MT (2001) Derivation of an eddy diffusivity depending on source distance for vertically inhomogeneous turbulence in a convective boundary layer. J Appl Meteorol 40(7):1233–1240

    Article  Google Scholar 

  • Demuth C (1978) A contribution to the analytical steady solution of the diffusion equation for line sources. Atmos Environ (1967) 12(5):1255–1258

    Article  Google Scholar 

  • Detto M, Montaldo N, Albertson JD, Mancini M, Katul G (2006) Soil moisture and vegetation controls on evapotranspiration in a heterogeneous Mediterranean ecosystem on Sardinia, Italy. Water Resour Res 42(8)

  • Dumortier P, de la Motte LG, Andriamandroso A, Aubinet M, Beckers Y, Bindelle J, De Cock N, Lebeau F, Heinesch B (2021) Beef cattle methane emission estimation using the eddy covariance technique in combination with geolocation. Agric For Meteorol 297(108):249

    Google Scholar 

  • Dyer A (1974) A review of flux-profile relationships. Bound Layer Meteorol 7(3):363–372

    Article  Google Scholar 

  • Eckman RM (1994) Re-examination of empirically derived formulas for horizontal diffusion from surface sources. Atmos Environ 28(2):265–272

    Article  Google Scholar 

  • Ema’a Ema’a JM, Ben-Bolie GH, Owono Ateba P (2013) A first-order WKB approximation for air pollutant dispersion equation. Atmos Res 120:162–169

    Google Scholar 

  • Fitzgerald GF, Trouton FT (1893) On the nomenclature for radiant energy. Nature 49(1259):149–149

    Article  Google Scholar 

  • Foken T (2006) 50 years of the Monin–Obukhov similarity theory. Bound Layer Meteorol 119(3):431–447

    Article  Google Scholar 

  • Gash J (1986) A note on estimating the effect of a limited fetch on micrometeorological evaporation measurements. Bound Layer Meteorol 35(4):409–413

    Article  Google Scholar 

  • Göckede M, Markkanen T, Mauder M, Arnold K, Leps JP, Foken T (2005) Validation of footprint models using natural tracer measurements from a field experiment. Agric For Meteorol 135(1–4):314–325

    Article  Google Scholar 

  • Göckede M, Foken T, Aubinet M, Aurela M, Banza J, Bernhofer C, Bonnefond JM, Brunet Y, Carrara A, Clement R et al (2008) Quality control of CarboEurope flux data-part 1: coupling footprint analyses with flux data quality assessment to evaluate sites in forest ecosystems. Biogeosciences 5(2):433–450

    Article  Google Scholar 

  • Goncalves GA, Quadro RS, Buske D (2013) An analytical formulation for pollutant dispersion simulation in the atmospheric boundary layer. J Environ Prot 4(08):57

    Article  Google Scholar 

  • Gryning SE, Van Ulden P, Larsen RE (1983) Dispersion from a continuous ground-level source investigated by a K model. Q J R Meteorol Soc 109(460):355–364

    Article  Google Scholar 

  • Gryning SE, Holtslag A, Irwin JS, Sivertsen B (1987) Applied dispersion modelling based on meteorological scaling parameters. Atmos Environ 21(1):79–89

    Article  Google Scholar 

  • Guerrero JP, Pimentel LCG, Oliveira-Júnior JFd, Heilbron Filho PFL, Ulke AG (2012) A unified analytical solution of the steady-state atmospheric diffusion equation. Atmos Environ 55:201–212

    Article  Google Scholar 

  • Haenel HD, Grünhage L (1999) Footprint analysis: a closed analytical solution based on height-dependent profiles of wind speed and eddy viscosity. Bound Layer Meteorol 93(3):395–409

    Article  Google Scholar 

  • Högström U (1988) Non-dimensional wind and temperature profiles in the atmospheric surface layer: a re-evaluation. Bound Layer Meteorol 42:55–78

    Article  Google Scholar 

  • Horst T (1999) The footprint for estimation of atmosphere-surface exchange fluxes by profile techniques. Bound Layer Meteorol 90(2):171–188

    Article  Google Scholar 

  • Horst T (2001) Comment on footprint analysis: a closed analytical solution based on height-dependent profiles of wind speed and eddy viscosity’. Bound layer meteorol 101(3):435–447

    Article  Google Scholar 

  • Horst T, Weil J (1992) Footprint estimation for scalar flux measurements in the atmospheric surface layer. Bound Layer Meteorol 59(3):279–296

    Article  Google Scholar 

  • Horst T, Weil J (1994) How far is far enough? The fetch requirements for micrometeorological measurement of surface fluxes. J Atmos Ocean Technol 11(4):1018–1025

    Article  Google Scholar 

  • Hsieh CI, Katul G, Chi TW (2000) An approximate analytical model for footprint estimation of scalar fluxes in thermally stratified atmospheric flows. Adv Water Resour 23(7):765–772

    Article  Google Scholar 

  • Huang C (1979) A theory of dispersion in turbulent shear flow. Atmos Environ (1967) 13(4):453–463

    Article  Google Scholar 

  • Karjanto N (2022) Perturbed potential temperature field in the atmospheric boundary layer. Z Agnew Math Mech 102(8):e202100484

    Google Scholar 

  • Kljun N, Calanca P, Rotach M, Schmid H (2004) A simple parameterisation for flux footprint predictions. Bound Layer Meteorol 112(3):503–523

    Article  Google Scholar 

  • Kljun N, Calanca P, Rotach M, Schmid HP (2015) A simple two-dimensional parameterisation for Flux Footprint Prediction (FFP). Geosci Model Dev 8(11):3695–3713

    Article  Google Scholar 

  • Kormann R, Meixner FX (2001) An analytical footprint model for non-neutral stratification. Bound Layer Meteorol 99(2):207–224

    Article  Google Scholar 

  • Krapez JC (2016) Heat diffusion in inhomogeneous graded media: chains of exact solutions by joint Property & Field Darboux Transformations. Int J Heat Mass Transf 99:485–503

    Article  Google Scholar 

  • Krapez JC (2018) The \({\rm sech}({\hat{ }})\)-type profiles: a Swiss-Army Knife for exact analytical modeling of thermal diffusion and wave propagation in graded media. Int J Thermophys 39(7):86

    Article  Google Scholar 

  • Krapez JC (2019) Linear, trigonometric and hyperbolic profiles of thermal effusivity in the Liouville space and related quadrupoles: Simple analytical tools for modeling graded layers and multilayers. Int J Thermal Sci 136:182–199

    Article  Google Scholar 

  • Krapez JC, Dohou E (2014) Thermal quadrupole approaches applied to improve heat transfer computations in multilayered materials with internal heat sources. Int J Thermal Sci 81:38–51

    Article  Google Scholar 

  • Kumar P, Sharan M (2010) An analytical model for dispersion of pollutants from a continuous source in the atmospheric boundary layer. Proc R Soc A Math Phys Eng Sci 466(2114):383–406

    Google Scholar 

  • Kumar P, Sharan M (2012) Parameterization of the eddy diffusivity in a dispersion model over homogeneous terrain in the atmospheric boundary layer. Atmos Res 106:30–43

    Article  Google Scholar 

  • Leclerc MY, Foken T (2014) Footprints in micrometeorology and ecology, vol 239. Springer

    Book  Google Scholar 

  • Leclerc M, Thurtell G (1990) Footprint prediction of scalar fluxes using a Markovian analysis. Bound Layer Meteorol 52(3):247–258

    Article  Google Scholar 

  • Loubet B, Génermont S, Ferrara R, Bedos C, Decuq C, Personne E, Fanucci O, Durand B, Rana G, Cellier P (2010) An inverse model to estimate ammonia emissions from fields. Eur J Soil Sci 61(5):793–805

    Article  Google Scholar 

  • Ma J, Daggupaty S (1998) A generalized analytical solution for turbulent dispersion with inhomogeneous wind and diffusion coefficient. Environ Modell Assess 3(4):239–248

    Article  Google Scholar 

  • Maillet D, Andre S, Batsale J, Degiovanni A, Moyne C (2000) Thermal quadrupoles: solving the heat equation through integral transforms. Wiley

    Google Scholar 

  • Mangia C, Moreira D, Schipa I, Degrazia G, Tirabassi T, Rizza U (2002) Evaluation of a new eddy diffusivity parameterisation from turbulent Eulerian spectra in different stability conditions. Atmos Environ 36(1):67–76

    Article  Google Scholar 

  • Moreira D, Vilhena M, Tirabassi T, Buske D, Cotta R (2005) Near-source atmospheric pollutant dispersion using the new GILTT method. Atmos Environ 39(34):6289–6294

    Article  Google Scholar 

  • Moreira DM, Tirabassi T, Carvalho JC (2005) Plume dispersion simulation in low wind conditions in stable and convective boundary layers. Atmos Environ 39(20):3643–3650

    Article  Google Scholar 

  • Moreira DM, Vilhena M, Carvalho JC, Degrazia GA (2005) Analytical solution of the advection–diffusion equation with nonlocal closure of the turbulent diffusion. Environ Modell Soft 20(10):1347–1351

    Article  Google Scholar 

  • Moreira DM, Vilhena MT, Tirabassi T, Costa C, Bodmann B (2006) Simulation of pollutant dispersion in the atmosphere by the Laplace transform: the ADMM approach. Water Air Soil Pollut 177(1):411–439

    Article  Google Scholar 

  • Moreira D, Vilhena M, Buske D, Tirabassi T (2009) The state-of-art of the GILTT method to simulate pollutant dispersion in the atmosphere. Atmos Res 92(1):1–17

    Article  Google Scholar 

  • Moreira DM, Tirabassi T, Vilhena M, Goulart A (2010) A multi-layer model for pollutant dispersion with dry deposition to the ground. Atmos Environ 44(15):1859–1865

    Article  Google Scholar 

  • Moreira DM, Vilhena MT, Tirabassi T, Buske D, da Costa CP (2010) Comparison between analytical models to simulate pollutant dispersion in the atmosphere. Int J Environ Waste Manag 6(3–4):327–344

    Article  Google Scholar 

  • Moreira DM, Moraes AC, Goulart AG, de Almeida Albuquerque TT (2014) A contribution to solve the atmospheric diffusion equation with eddy diffusivity depending on source distance. Atmos Environ 83:254–259

    Article  Google Scholar 

  • Nascimento EGS, Moreira DM, de Almeida Albuquerque TT (2017) The development of a new model to simulate the dispersion of rocket exhaust clouds. Aerosp Sci Technol 69:298–312

    Article  Google Scholar 

  • Nieuwstadt F, Van Ulden A (1978) A numerical study on the vertical dispersion of passive contaminants from a continuous source in the atmospheric surface layer. Atmos Environ 12(11):2119–2124

    Article  Google Scholar 

  • Optis M, Monahan A, Bosveld FC (2016) Limitations and breakdown of Monino-Obukhov similarity theory for wind profile extrapolation under stable stratification. Wind Energy 19(6):1053–1072

    Article  Google Scholar 

  • Pasquill F, Smith F (1983) Atmospheric diffusion. Wiley

    Google Scholar 

  • Paulson CA (1970) The mathematical representation of wind speed and temperature profiles in the unstable atmospheric surface layer. J Appl Meteorol 9(6):857–861

    Article  Google Scholar 

  • Rannik Ü, Sogachev A, Foken T, Göckede M, Kljun N, Leclerc MY, Vesala T (2012) Footprint analysis. In: Eddy Covariance, pp 211–261. Springer

  • Rey-Sanchez C, Arias-Ortiz A, Kasak K, Chu H, Szutu D, Verfaillie J, Baldocchi D (2022) Detecting hot spots of methane flux using footprint-weighted flux maps. J Geophys Res Biogeosci 127(8):e2022JG006977

    Article  Google Scholar 

  • Schmid H (1994) Source areas for scalars and scalar fluxes. Bound Layer Meteorol 67(3):293–318

    Article  Google Scholar 

  • Schmid HP (2002) Footprint modeling for vegetation atmosphere exchange studies: a review and perspective. Agric For Meteorol 113(1–4):159–183

    Article  Google Scholar 

  • Schmid H, Oke T (1990) A model to estimate the source area contributing to turbulent exchange in the surface layer over patchy terrain. Q J R Meteorol Soc 116(494):965–988

    Article  Google Scholar 

  • Schuepp P, Leclerc M, MacPherson J, Desjardins R (1990) Footprint prediction of scalar fluxes from analytical solutions of the diffusion equation. Bound Layer Meteorol 50(1–4):355–373

    Article  Google Scholar 

  • Stull RB (2012) An introduction to boundary layer meteorology, vol 13. Springer

    Google Scholar 

  • Timmermans W, Su Z, Olioso A (2009) Footprint issues in scintillometry over heterogeneous landscapes. Hydrol Earth Syst Sci 13(11):2179–2190

    Article  Google Scholar 

  • Tirabassi T, Tagliazucca M, Zannetti P (1986) KAPPA-G, a non-Gaussian plume dispersion model: description and evaluation against tracer measurements. J Air Pollut Control Assoc 36(5):592–596

    Article  Google Scholar 

  • Toutain J, Battaglia J, Pradere C, Pailhes J, Kusiak A, Aregba W, Batsale J (2011) Numerical inversion of Laplace transform for time resolved thermal characterization experiment. J Heat Transf 133(4):044504

    Article  Google Scholar 

  • Van Ulden A (1978) Simple estimates for vertical diffusion from sources near the ground. Atmos Environ 12(11):2125–2129

    Article  Google Scholar 

  • Venkatram A, Snyder MG, Heist DK, Perry SG, Petersen WB, Isakov V (2013) Re-formulation of plume spread for near-surface dispersion. Atmos Environ 77:846–855

    Article  Google Scholar 

  • Vesala T, Kljun N, Rannik Ü, Rinne J, Sogachev A, Markkanen T, Sabelfeld K, Foken T, Leclerc M (2008) Flux and concentration footprint modelling: state of the art. Environ Pollut 152(3):653–666

    Article  Google Scholar 

  • Vilhena M, Rizza U, Degrazia GA, Mangia C, Moreira D, Tirabassi T (1998) An analytical air pollution model: development and evaluation. Contrib Atmos Phys 71(3):315–320

    Google Scholar 

  • Wang Q, Zhan H (2015) On different numerical inverse Laplace methods for solute transport problems. Adv Water Resour 75:80–92

    Article  Google Scholar 

  • Wilson JD (2015) Computing the flux footprint. Bound Layer Meteorol 156(1):1–14

    Article  Google Scholar 

  • Wortmann S, Vilhena MT, Moreira D, Buske D (2005) A new analytical approach to simulate the pollutant dispersion in the PBL. Atmos Environ 39(12):2171–2178

    Article  Google Scholar 

  • Yeh GT (1975) Green’s functions of a diffusion equation. Geophys Res Lett 2(7):293–296

    Article  Google Scholar 

  • Yeh GT, Huang CH (1975) Three-dimensional air pollutant modeling in the lower atmosphere. Bound Layer Meteorol 9(4):381–390

    Article  Google Scholar 

Download references

Acknowledgements

The authors are indebted to Claire Sarrat, Clélia Robert and Jean-Martial Cohard for fruitful discussions and to the anonymous reviewers for their suggestions for improving this paper.

Author information

Authors and Affiliations

Authors

Contributions

J-CK and GAK worked on the conceptualization (Liouville transformation), GAK performed preliminary profile fittings, and J-CK developed the methodology (semi-analytical model and surrogate models). J-CK wrote the manuscript text, which was reviewed by GAK. The project was supervised by J-CK.

Corresponding author

Correspondence to J.-C. Krapez.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Below is the link to the electronic supplementary material.

Supplementary file 1 (pdf 4061 KB)

Appendices

Appendix 1: Power-Law Profiles

We discuss the Diffusion-Ascent-Associated Advection Distance (DAAAD), \(\xi ^2\), and the aerodynamic inertivity b as functions of height z, for the two simplest cases: constant profiles and PL profiles. Assuming constant wind speed \(u(z)=u_1\), constant eddy diffusivity \(K(z)=K_1\), and a line source of strength Q at height \(z_b=0\), the Liouville scale defined in Eq. 17, \(\xi (z)\), which can be seen as the integrated (square root of) downwind advection distance of a cloud of particles while vertically diffusing up to height z, is in power 1 of height: \(\xi (z)=\sqrt{u_1/K_1}z\). The DAAAD \(\xi (z)^2\) is thus in power 2 of height. The aerodynamic inertivity of the atmosphere is constant: \(b(z)=b_1=\)\(\sqrt{u_1K_1}\). The scalar field follows the classical Gaussian plume model:

$$\begin{aligned} \chi (x,z)=\frac{Q}{\sqrt{\pi K_1 u_1 x}}\exp \left( -\frac{u_1 z^2}{4K_1 x}\right) . \end{aligned}$$
(44)

The concentration at a given height z rises to a maximum at a downwind distance \(x_{\chi max}(z)=\frac{u_1 z^2}{2K_1}\) and then decays continuously as distance x is further increased. Hence, the DAAAD \(\xi (z)^2\) corresponds to two times \(x_{\chi max}(z)\). At this distance the concentration has dropped by about 9.2% from the maximum concentration observed at height z.

When expressed in terms of aerodynamic inertivity and \(\xi \), the 2D concentration field is:

$$\begin{aligned} \chi (x,\xi )=\frac{Q}{b_1\sqrt{\pi x}}\exp \left( -\frac{\xi ^2}{4x}\right) , \end{aligned}$$
(45)

where it can be seen that concentration is inversely proportional to the local aerodynamic inertivity \(b(z)=b_1\), which motivated the choice of this name for this parameter.

Let us now assume the power-law profiles in Eq. 5 with again a line source at height \(z_b=0\). The Liouville scale \(\xi (z)\) and the aerodynamic inertivity b(z) are themselves PL functions of height:

$$\begin{aligned} \xi (z)= & {} \xi _1\left( \frac{z}{z_1}\right) ^{r/2}\quad ;\quad \xi _1 \equiv z_1\frac{2}{r}\sqrt{\frac{u_1}{K_1}}\quad ;\quad r \equiv m-n+2, \end{aligned}$$
(46)
$$\begin{aligned} b(z)= & {} b_1\left( \frac{z}{z_1}\right) ^{\frac{m+n}{2}} \quad ;\quad b_1 \equiv \sqrt{u_1 K_1}. \end{aligned}$$
(47)

As a consequence, \(\sqrt{b}\) is itself a PL function of \(\xi \) with power \(\mu -1/2\) where \(\mu \equiv (m+1)/(m-n+2)\). Notice that it is therefore a particular case of the \(\sqrt{b}\)-EPL profiles in Eq. 27 with \(\nu =\mu -1\). The scalar field (Robert’s solution) expresses as:

$$\begin{aligned} \chi (x,z)=\frac{rQ}{z_1 u_1 \varGamma (\mu )}\left( \frac{u_1 z_1^2}{K_1 r^2 x}\right) ^{\mu }\exp \left( -\frac{u_1 z_1^2}{K_1 r^2 x}\left( \frac{z}{z_1}\right) ^r\right) . \end{aligned}$$
(48)

Upon introducing \(\xi (z)\) expressed in Eq. 46 we obtain Eq. 7. This expression corresponds to Eq. 20 in Kormann and Meixner (2001), knowing their “\(\xi \)” variable is equivalent to present \(\xi ^2/4\). Further introducing the aerodynamic inertivity b(z) expressed in Eq. 47 we obtain:

$$\begin{aligned} \chi (x,\xi )=\frac{2Q}{b \xi \varGamma (\mu )} \left( \frac{\xi ^2}{4x}\right) ^{\mu } \exp \left( -\frac{\xi ^2}{4x}\right) , \end{aligned}$$
(49)

which shows again that the concentration is inversely proportional to the local aerodynamic inertivity \(b(\xi (z))\). The inferred concentration footprint can be nondimensionalized by multiplying it by \(b \xi \). This gives:

$$\begin{aligned} \left\langle \overline{f_{\chi }}^y\right\rangle \equiv \overline{f_{\chi }}^y{b \xi }=\frac{2}{\varGamma (\mu )} \left( \frac{\xi ^2}{4x}\right) ^{\mu } \exp \left( -\frac{\xi ^2}{4x}\right) , \end{aligned}$$
(50)

It is easy to see that the DAAAD \(\xi (z)^2\) and the downwind distance \(x_{\chi max}(z)\) at which the concentration footprint reaches its maximum are now related through \(\xi (z)^2=4\mu x_{\chi max}(z)\) which implies that \(x_{\chi max}(z)\) increases to the power r of height. Similarly, the flux footprint in Eq. 11 can be nondimensionalized by multiplying it by \(\xi ^2\), which gives:

$$\begin{aligned} \left\langle \overline{f_{\varphi }}^y\right\rangle \equiv \overline{f_{\varphi }}^y{\xi ^2 }=\frac{4}{\varGamma (\mu )}\left( \frac{\xi ^2}{4x}\right) ^{\mu +1} \exp \left( -\frac{\xi ^2}{4x}\right) . \end{aligned}$$
(51)

The DAAAD \(\xi (z)^2\) and the downwind distance \(x_{\varphi max}(z)\) at which the flux footprint reaches its maximum are related through \(\xi (z)^2=4(\mu +1) x_{\varphi max}(z)\). The downwind distances at which the concentration and flux footprints reach their maximum are in a constant ratio \((\mu +1)/\mu \).

The inverse Gamma distribution is defined by:

$$\begin{aligned} f\left( x,\mu ,\beta \right) =\frac{1}{\beta \varGamma (\mu )}\left( \frac{\beta }{x}\right) ^{\mu +1} \exp \left( -\frac{\beta }{x}\right) , \end{aligned}$$
(52)

with \(\mu >0\), the shape parameter. The flux footprint in Eq. 11 can thus be expressed vs the inverse Gamma distribution as:

$$\begin{aligned} \overline{f_{\varphi }}^y=f\left( x,\mu ,\frac{\xi ^2}{4}\right) , \end{aligned}$$
(53)

and the cumulative flux footprint can be expressed vs the regularized Gamma function \(F\left( x,\mu ,\beta \right) \) as:

$$\begin{aligned} \overline{F_{\varphi }}^y=F\left( x,\mu ,\frac{\xi ^2}{4}\right) \equiv \frac{\varGamma \left( \mu ,\frac{\xi ^2}{4x}\right) }{\varGamma (\mu )}, \end{aligned}$$
(54)

where \(\varGamma \left( \mu ,t\right) \) is the upper incomplete Gamma function defined as \(\varGamma \left( \mu ,t\right) =\int _{t}^{\infty }w^{\mu -1}e^{-w}dw\). The limit of the cumulative flux footprint for \(x \rightarrow \infty \), is 1, as expected. The condition \(\mu >0\) must be fullfilled, which is the case for PL profiles fitting realistic atmospheric profiles.

Similarly, the concentration footprint can be expressed vs the inverse Gamma distribution as:

$$\begin{aligned} \overline{f_{\chi }}^y=\frac{\xi }{2b}\frac{\varGamma (\mu -1)}{\varGamma (\mu )}f\left( x,\mu -1,\frac{\xi ^2}{4}\right) , \end{aligned}$$
(55)

and the cumulative concentration footprint can be expressed vs the regularized Gamma function \(F\left( x,\mu ,\beta \right) \) as:

$$\begin{aligned} \overline{F_{\chi }}^y=\frac{\xi }{2b}\frac{\varGamma (\mu -1)}{\varGamma (\mu )}F\left( x,\mu -1,\frac{\xi ^2}{4}\right) = \frac{\xi }{2b}\frac{\varGamma \left( \mu -1,\frac{\xi ^2}{4x}\right) }{\varGamma (\mu )}. \end{aligned}$$
(56)

The limit of the cumulative concentration footprint for \(x \rightarrow \infty \) is \(\frac{\xi }{2b}\frac{\varGamma (\mu -1)}{\varGamma (\mu )}\) provided \(\mu >1\), otherwise the function is unbounded.

Appendix 2: Quadrupole Method

The four entries ABCD of the quadrupole related to a graded layer of \(\sqrt{b}\)-EPL type, Eqs. 27 and 28, are expressed as:

$$\begin{aligned} A= & {} \frac{{{\check{\xi }}}_{1}h_{1}\sqrt{p}}{h_{0}}\left( I_{\nu ,0}K_{\nu +1,1}+I_{\nu +1,1}K_{\nu ,0}\right) -\frac{2\nu A_{D}{{\check{\xi }}}_{1}^{-\nu }}{h_{0}} \left( I_{\nu ,0}K_{\nu ,1}-I_{\nu ,1}K_{\nu ,0}\right) , \end{aligned}$$
(57)
$$\begin{aligned} B= & {} -\frac{1}{h_{0}h_{1}}\left( I_{\nu ,0}K_{\nu ,1}-I_{\nu ,1}K_{\nu ,0}\right) , \end{aligned}$$
(58)
$$\begin{aligned}{} & {} \begin{aligned} C&= -h_{0}h_{1}{{\check{\xi }}}_{0}{{\check{\xi }}}_{1}p \left( I_{\nu +1,0}K_{\nu +1,1}-I_{\nu +1,1}K_{\nu +1,0}\right) \\&\quad +\, 2\nu A_{D}h_{0}{{\check{\xi }}}_{0}{{\check{\xi }}}_{1}^{-\nu }\sqrt{p} \left( I_{\nu +1,0}K_{\nu ,1}+I_{\nu ,1}K_{\nu +1,0}\right) \\&\quad -\, 2\nu A_{D}h_{1}{{\check{\xi }}}_{0}^{-\nu }{{\check{\xi }}}_{1}\sqrt{p} \left( I_{\nu ,0}K_{\nu +1,1}+I_{\nu +1,1}K_{\nu ,0}\right) \\&\quad +\, 4\nu ^2 A_{D}^2{{\check{\xi }}}_{0}^{-\nu }{{\check{\xi }}}_{1}^{-\nu } \left( I_{\nu ,0}K_{\nu ,1}-I_{\nu ,1}K_{\nu ,0}\right) , \end{aligned} \end{aligned}$$
(59)
$$\begin{aligned} D= & {} \frac{{{\check{\xi }}}_{0}h_{0}\sqrt{p}}{h_{1}}\left( I_{\nu +1,0}K_{\nu ,1}+I_{\nu ,1}K_{\nu +1,0}\right) +\frac{2\nu A_{D}{{\check{\xi }}}_{0}^{-\nu }}{h_{1}} \left( I_{\nu ,0}K_{\nu ,1}-I_{\nu ,1}K_{\nu ,0}\right) . \end{aligned}$$
(60)

In the brackets present in Eqs. 5760, the symbols \(I_{\lambda ,k}\) and \(K_{\lambda ,k}\) have been introduced, where \(\lambda \) takes the values \(\nu \) or \(\nu +1\) and k takes the values 0 or 1. They represent respectively \(I_{\lambda }(\sqrt{p}{{\check{\xi }}}_{k})\) and \(K_{\lambda }(\sqrt{p}{{\check{\xi }}}_{k})\). Furthermore, \({{\check{\xi }}}\) and h with indices 0 and 1 refer to the values taken by these functions (see Eqs. 2627) at the low, resp. high boundary of the considered layer.

The bottom layer, whose lower end is fixed at a height correponding to the roughness length, \(z=z_0\), requires particular attention (notice that \(z_b\) in the definition of the Liouville scale in Eq. 17 is then set to \(z_0\)). As a matter of fact, wind speed vanishes at \(z=z_0\), hence the aerodynamic inertivity vanishes as well. The \(\sqrt{b}\)-EPL profile in Eq. 27 should thus be chosen so that \(b(\xi )\) vanishes at \(\xi =0\). Setting the constant \(\xi _{c}\) to 0 allows achieving this. One should then keep only those power-functions (from the two) that have a positive exponent (\(1/2+\nu \) and \(1/2-\nu \)). Assuming that the MO similarity theory is valid down to \(z_0\), the effusivity profile follows asymptotically the neutral profile, i.e. \(b(z)\propto z ln(z/z_0)\). While fitting a \(\sqrt{b}\)-EPL profile to the neutral profile (after expressing it in terms of \(\xi \)) we checked that both power-law functions of exponent \(1/2+\nu \) and \(1/2-\nu \) in Eq. 27 can be kept (indeed, the optimal value for \(\nu \) is about 0.3). As a consequence, we have \(h_{0}=0\) and \(h_{1}=A_{B}\xi _{1}^{\nu }+A_{D}\xi _{1}^{-\nu }\) for the bottom layer. However, several terms in Eqs. 5760 are now unbounded because involving \(\check{\xi _0}=\xi _0=0\), in particular the terms \(\check{\xi _0}^{-\nu }\) and \(K_{\nu ,0}\) and \(K_{\nu +1,0}\) which correspond here to \(K_{\nu }(0)\) and \(K_{\nu +1}(0)\). Expansion in power series of the Bessel functions is required to solve the singularities. Notice that only the two entries C and D prove to be necessary when concentration and flux are to be calculated at heights not lower than the top of the bottom layer, which will be the case (refer to Eq. 34). After some manipulations (see the Supplementary Material Sect. S1–2) we find the following expressions for the entries C and D of the bottom quadrupole:

$$\begin{aligned}{} & {} \begin{aligned} C&=A_{B}h_{1}{{\check{\xi }}}_{1}\sqrt{p}^{\;-\nu +1}2^{\nu }\varGamma (\nu +1)I_{\nu +1,1}\\&\quad +\, \nu A_{B}A_{D}{{\check{\xi }}}_{1}^{-\nu }\sqrt{p}^{\;-\nu }2^{\nu +1}\varGamma (\nu +1)I_{\nu ,1}\\&\quad -\, \nu A_{D}h_{1}{{\check{\xi }}}_{1}\sqrt{p}^{\nu +1} \left[ 2^{-\nu +1}/\varGamma (\nu +1)K_{\nu +1,1}+2^{-\nu }\varGamma (-\nu )I_{\nu +1,1}\right] \\&\quad +\, \nu ^2 A_{D}^2{{\check{\xi }}}_{1}^{-\nu }\sqrt{p}^{\nu } \left[ 2^{-\nu +2}/\varGamma (\nu +1)K_{\nu ,1}-2^{-\nu +1}\varGamma (-\nu )I_{\nu ,1}\right] , \end{aligned} \end{aligned}$$
(61)
$$\begin{aligned}{} & {} \begin{aligned}&D=A_{B}h_{1}^{-1}\sqrt{p}^{\;-\nu }2^{\nu }\varGamma \left( \nu +1 \right) I_{\nu ,1}\\&\quad \quad +\nu A_{D}h_{1}^{-1}\sqrt{p}^{\nu }\left[ 2^{-\nu +1}/\varGamma \left( \nu +1\right) K_{\nu ,1}-2^{-\nu }\varGamma \left( -\nu \right) I_{\nu ,1}\right] . \end{aligned} \end{aligned}$$
(62)

We now develop the expression of the admittance at the bottom of a semi-infinite layer having a \(\sqrt{b}\)-EPL profile (third option for the closure of the multilayer problem discussed in Sect. 2.5). From the definition of the admittance we obtain, for a generally graded layer (Krapez 2016, Eq. 19):

$$\begin{aligned} Y=\sqrt{b}\left( \sqrt{b}'\psi -\sqrt{b}\psi '\right) \psi ^{-1}. \end{aligned}$$
(63)

On the other hand, from Eqs. 2327 and 28, the general form for the concentration is:

$$\begin{aligned} X(p,\xi )=\frac{1}{h({\check{\xi }})}\left[ A_I I_{\nu }(\sqrt{p}{\check{\xi }})+A_K K_{\nu }(\sqrt{p}{\check{\xi }})\right] . \end{aligned}$$
(64)

The asymptotic expansions of the modified Bessel functions for \(\vert t\vert \rightarrow +\infty \) are:

$$\begin{aligned} I_{\nu }(t)\approx \frac{1}{\sqrt{2\pi t}} \exp {\left( t\right) }\quad ;\quad K_{\nu }(t)\approx \sqrt{\frac{\pi }{2t}} \exp {\left( -t\right) }, \end{aligned}$$
(65)

which shows that the particular solution involving \(I_{\nu }\) in Eq. 64 is unbounded for \(\xi \rightarrow \infty \) and should thus be discarded. The admittance thus simplifies as:

$$\begin{aligned} Y =\frac{h{{\check{\xi }}}\left( h'K_{\nu }-hK_{\nu }'\right) }{K_{\nu }}. \end{aligned}$$
(66)

The admittance at \(\xi =\xi _0\), the bottom of the topmost (“sacrificial”) layer, finally reads:

$$\begin{aligned} Y_{top}=h_{0}\left( h_{0}{{\check{\xi }}}_{0}\sqrt{p}\frac{K_{\nu +1,0}}{K_{\nu ,0}} -2\nu A_{D}{{\check{\xi }}}_{0}^{-\nu }\right) . \end{aligned}$$
(67)

Appendix 3: Verification of the Semi-analytical Model

Matlab R2021a on a PC equipped with an Intel core Xeon W-2123 @ 3.6GHz CPU and 8GB RAM has been employed for the computations and evaluation of the computation time. The first verification test was performed with the classical uK-PL(z) profiles (the expressions of related concentration and flux are in Appendix 1, Eqs. 50 and 51). The whole process, from step 1 to step 8, was applied (see Sect. 2.8). Notice that since any pair of uK-PL(z) profiles leads, in the Liouville space, to a b-PL\((\xi )\) profile, this one being a special case of the \(\sqrt{b}\)-EPL\(({\check{\xi }})\) profiles (see Appendix 1), the fitting in step 3 was thus formally trivial. In practice however, small (negligible) errors were observed at the end of step 4, which actually came from the quadrature computation to evaluate \(\xi (z)\) in step 2. In the end, the comparison with the analytical solution in Eq. 51 was very satisfactory, since the relative error was less than \(10^{-9}\).

We didn’t find in the literature other expressions for graded profiles u(z) and K(z) that are analytically solvable, namely leading to an exact and closed form analytical solution of the stationary advection–diffusion problem. Nevertheless, the “thermal” solutions developed in Krapez (2016) can offer a starting point to build such solvable profiles. Among the numerous possibilities, we selected the following inertivity profiles, which were called \(\tanh ^2\)-type profiles:

$$\begin{aligned} b(\xi )=b_{\infty }\left( \tanh \left( \xi /\xi _{c}\right) \right) ^2, \end{aligned}$$
(68)

where \(b_{\infty }\) is the asymptotic inertivity for increasing height and \(\xi _{c}\) a scale for the Liouville variable. Among the infinite number of pairs of profiles u(z) and K(z) from which these inertivity profiles \(b(\xi )\) can derive, we will consider, by way of illustration, the simplest one, i.e. the one embodying the same spatial variability in u(z) and K(z), thus ensuring a proportionality between z and \(\xi \):

$$\begin{aligned}&u(z)=u_{\infty }\left( \tanh \left( \left( z-z_0\right) /z_{c}\right) \right) ^2, \end{aligned}$$
(69a)
$$\begin{aligned}&K(z)=K_{\infty }\left( \tanh \left( \left( z-z_0\right) /z_{c}\right) \right) ^2, \end{aligned}$$
(69b)

where \(u_{\infty }\) and \(K_{\infty }\) are the limiting values of wind speed and diffusivity for increasing height and \(z_{c}\) is a scale for height. In this case \(b_{\infty }=\sqrt{u_{\infty }K_{\infty }}\), \(\xi =(z-z_0)\sqrt{u_{\infty }/K_{\infty }}\), and \(\xi _{c}=z_{c}\sqrt{u_{\infty }/K_{\infty }}\). As a result of a source of strength Q at \(z=z_{0}\) (i.e. \(\xi =0\)), the concentration is:

$$\begin{aligned} \chi (x,\xi )=\frac{Q}{b_{\infty }\sqrt{\pi x}}\left( 1+\frac{\xi \xi _{c}}{2x\tanh \left( \xi /\xi _{c}\right) }\right) \exp \left( -\frac{\xi ^2}{4x}\right) , \end{aligned}$$
(70)

and the flux is:

$$\begin{aligned} \varphi (x,\xi )=\frac{4Q}{\sqrt{\pi }\xi ^2} \left[ \left( 1-\frac{\xi _{c}}{\xi }\sqrt{\frac{b}{b_{\infty }}}\right) \gamma ^{3/2}+ 2\frac{\xi _{c}}{\xi }\sqrt{\frac{b}{b_{\infty }}}\gamma ^{5/2} \right] \exp (-\gamma ), \end{aligned}$$
(71)

with \(\gamma =\xi ^2/(4x)\) (see Supplementary material, Sect. S2). Even though the \(\tanh ^2\)-type profiles in Eq. 69 are quite far from MOST profiles, they present the mathematical advantage of leading to simple, closed-form analytical expressions of the footprints; in addition, compared to uK-PL(z) profiles, they present the characteristic of being bounded. Hence, they have an interest for benchmarking of semi-analytical or numerical solutions.

The semi-analytical model was tested with the \(\tanh ^2\)-type profiles in Eq. 69 by comparing the computed flux to the theoretical values obtained from Eq. 71. Starting with one EPL layer on each side of the assumed measurement level, the layers were subdivided until the fitting (relative) error was everywhere lower than a prescribed value ranging from 0.7 down to \(3\times 10^{-5}\). The number of required layers ranged from 3 to 37. The maximum (relative) error with respect to the theoretical flux in Eq. 71 decreased from 0.07 to \(10^{-6}\) (the relative error was obtained by dividing the difference in footprint values by the maximum of the theoretical footprint). Discretizing in 11 layers was enough to reach a relative error lower than about \(2\times 10^{-3}\) in the aerodynamic inertivity and \(10^{-4}\) in the flux-footprint. In this case, the CPU time was about 0.66 s (0.6 s for the aerodynamic-inertivity profile definition and fitting - step 1 to 4, and 0.06 s for the quadrupole manipulations and Laplace inversion for 100 points along x-axis—step 5 to 8). More information on the verification task, in particular a detailed analysis on the interplay between the fitting-error constraint, the number of discretized layers, and the footprint error can be found in the Supplementary Material Sect. S2.

Appendix 4: Monin–Obukhov Profiles

Over a horizontally homogeneous surface in steady-state conditions the vertical fluxes are nearly constant with height in the lower part of the boundary layer. The surface layer is commonly defined to be the layer where the vertical fluxes vary by less than 10% of their magnitude (Stull 2012). Among them are the momentum flux (\(-\overline{w'u'}\)) the sensible heat flux (\(-\overline{w'\theta '}\)) and the trace gas flux (\(-\overline{w'\chi '}\)) (w is the vertical wind speed, \(\theta \) the temperature, the overbar denotes Reynolds average and each primed variable represents the deviation from its average). This allows defining in the surface layer scales for velocity (friction velocity \(u_*=\sqrt{|\overline{w'u'}|}\)), temperature (\(\theta _*=-\overline{w'\theta '}/u_*\)) and tracer-concentration (\(\chi _*=-\overline{w'\chi '}/u_*\)). The Obukhov length is defined as:

$$\begin{aligned} L=-\frac{\theta _v u_*^3}{\kappa g \overline{w'\theta _v'}}, \end{aligned}$$
(72)

where \(\theta _v\) and \(\overline{w'\theta _v'}\) are the virtual potential temperature and buoyancy flux, respectively, g is the gravitational acceleration and \(\kappa \) is the von Karman constant (commonly accepted value is \(\kappa =0.4\)). The Obukhov length can be used to classify atmospheric stability conditions: L is negative when buoyancy flux is positive, namely upwards (unstable stratification), L is positive when when buoyancy flux is negative (stable), \(\vert L\vert \rightarrow \infty \) when buoyancy flux vanishes (neutral). The MO theory states that when properly scaled with \(u_*\), \(\theta _*\) and \(\chi _*\), the wind-shear, temperature and tracer-concentration gradients are functions (namely universal similarity functions) that depend only on the stability parameter \(\zeta \equiv z/L\), where L is the Obukhov length scale:

$$\begin{aligned}&\frac{\kappa z}{u_*}\frac{du}{dz}=\phi _m(\zeta ), \end{aligned}$$
(73a)
$$\begin{aligned}&\frac{\kappa z}{\theta _*}\frac{d\theta }{dz}=Pr_n\cdot \phi _{h}(\zeta ), \end{aligned}$$
(73b)
$$\begin{aligned}&\frac{\kappa z}{\chi _*}\frac{d\chi }{dz}=Sc_n\cdot \phi _{\chi }(\zeta ). \end{aligned}$$
(73c)

In this representation \(\phi _{h,\chi }\rightarrow 1\) for \(|\zeta |\rightarrow 0\) (neutral limit) whereas \(Pr_n\) and \(Sc_n\) are the turbulent Prandtl and Schmidt numbers in the neutral limit, respectively (see e.g. Leclerc and Foken (2014), Wilson (2015)).

The wind profile u(z) is obtained by integrating Eq. 73a which gives Eq. 38. On the other hand, upon applying the K-theory which expresses the turbulent flux \(\overline{w'\chi '}\) as \(-K\) times the gradient of concentration, where K is the eddy diffusivity related to tracer concentration, Eq. 73c leads to the expression of the eddy diffusivity in Eq. 39.

We considered the universal similarity functions initially proposed by Businger et al. (1971) and later revised by Högström (1988) (BH parameterizations) while taking \(\kappa =0.4\).

For the unstable case (\(L<0\)):

$$\begin{aligned}&\phi _m(\zeta )=\left( 1-19.3\zeta \right) ^{-1/4}, \end{aligned}$$
(74a)
$$\begin{aligned}&Pr_n\cdot \phi _h(\zeta )=Pr_n\cdot \left( 1-11.6\zeta \right) ^{-1/2}, \end{aligned}$$
(74b)

while, for the stable case (\(L>0\)):

$$\begin{aligned}&\phi _m(\zeta )=1+6\zeta , \end{aligned}$$
(75a)
$$\begin{aligned}&Pr_n\cdot \phi _h(\zeta )=Pr_n\cdot \left( 1+8.21\zeta \right) , \end{aligned}$$
(75b)

with \(Pr_n=0.95\). For the similarity function related to tracer concentration, we assumed both \(\phi _{\chi }(\zeta )=\phi _h(\zeta )\) and \(Sc_n=Pr_n=0.95\) although other values for \(Sc_n\), the neutral turbulent Schmidt number, could be considered as well.

Based on the expressions in Eqs. 74a and 75a, the integral in Eq. 38 takes the form proposed by Paulson (1970):

$$\begin{aligned} u(z)=\frac{u_*}{\kappa } \left[ \ln \left( \frac{z}{z_0}\right) -\psi _m(\zeta )+\psi _m(\zeta _0)\right] , \end{aligned}$$
(76)

where \(\zeta _0\equiv z_0/L\) and the integrated similarity function \(\psi _m\) is given by, if \(L<0\):

$$\begin{aligned} \psi _m(\zeta )= 2\ln \left( \frac{1+\phi _m^{-1}}{2}\right) +\ln \left( \frac{1+\phi _m^{-2}}{2}\right) -2\arctan (\phi _m^{-1})+\frac{\pi }{2}, \end{aligned}$$
(77)

where \(\phi _m(\zeta )\) is taken from Eq. 74a, while, if \(L>0\):

$$\begin{aligned} \psi _m(\zeta )=-6\zeta . \end{aligned}$$
(78)

In case the individual roughness elements are packed very closely together (e.g. forest canopy with close trees or cities with houses packed close together) a displacement distance d has to be introduced which is fraction of the mean height of the elements. The absolute height z appearing in Eqs. 38, 39, 74a78 should then be replaced by the relative height \(z-d\).

Typical profiles of the wind speed and eddy diffusivity according to MOST while taking into account the parameterisations shown before have been plotted in Fig. 4 in Supplementary Material Sect. S3.

Appendix 5: Parameterization of the Surrogate Model for the Flux Footprint

The surrogate model for the (crosswind-integrated) flux footprint is described in Eq. 40. The two parameters \(\mu _s\) and \(\xi _s^2\) on which this model is based were then fitted with respect to the scaled height \(z_m/z_0\) and and the stability parameter \(z_0/L\) (a scaling of \(\xi _s^2\) with \(\sqrt{z_0}\) was chosen to provide dimensionless results; in Eq. 40, the upwind distance x will thus have to be identically scaled).

We propose the following parameterizations, which, for ease, are given in reference to the neutral case (with a superscript “\(\textit{n}\)”) according to:

$$\begin{aligned}&\mu _s\left( \frac{z_0}{L},\frac{z_m}{z_0}\right) =\mu _s^{n} \cdot R_{\mu }\left( \frac{z_0}{L},\frac{z_m}{L}\right) , \end{aligned}$$
(79a)
$$\begin{aligned}&\frac{\xi _s}{\sqrt{z_0}}\left( \frac{z_0}{L},\frac{z_m}{z_0}\right) =\frac{\xi _s^{n}}{\sqrt{z_0}} \cdot R_{\xi } \left( \frac{z_0}{L},\frac{z_m}{L}\right) . \end{aligned}$$
(79b)

The parameters for the neutral case are approximated by:

$$\begin{aligned}&\mu _s^n=0.7577+0.2423 \cdot \left( \tanh \left( 1.217\cdot X^{0.4135} \right) \right) ^{16.38}, \end{aligned}$$
(80a)
$$\begin{aligned}&\frac{\xi _s^n}{\sqrt{z_0}}=1.735 \cdot X \cdot Y^p, \end{aligned}$$
(80b)

where

$$\begin{aligned} \begin{aligned}&p=0.5164-6.604\times 10^{-3}\ln (Y)+2.578\times 10^{-4}\ln (Y)^2,\\&X=\ln \left( \frac{z_m}{z_0}\right) ,\\&Y=\frac{z_m}{z_0}-1. \end{aligned} \end{aligned}$$
(81)

The functions \(R_{\mu }\) and \(R_{\xi }\) were parameterized as described in the following expressions with possibly different forms depending on whether the atmosphere is stable or unstable (subscript “-” for unstable and “+” for stable):

$$\begin{aligned} R_{\mu }^{-}=&1+A_1 \tanh \left[ A_2\ln \left( A_3 Z_{\mu }+1\right) \right] , \end{aligned}$$
(82a)
$$\begin{aligned} R_{\mu }^{+}=&\left\{ 1+A_1 \tanh \left[ A_2\ln \left( A_3 Z_{\mu }+1\right) \right] \exp \left( A_4 Z_{\mu }^{1/4} \right) \right\} ^{-1}, \end{aligned}$$
(82b)
$$\begin{aligned} R_{\xi }^{-},\;R_{\xi }^{+}=&\left( 1+A_1 Z_{\xi }^{A_2}\right) ^{A_3}, \end{aligned}$$
(82c)

with

$$\begin{aligned}&Z_{\mu }= \frac{z_0}{\left| L\right| } \left( \frac{z_m}{z_0} -1 \right) , \end{aligned}$$
(83a)
$$\begin{aligned}&Z_{\xi }=\frac{z_m}{\left| L\right| }. \end{aligned}$$
(83b)

The coefficients \(A_1\) to \(A_5\) are expressed as functions of the parameter \(z_0/L\) of the following type:

$$\begin{aligned} A_{i}=B_1+B_2 \left| \frac{z_0}{L} \right| ^{B_3},\quad i=1 \ldots 5, \end{aligned}$$
(84)

where the coefficients \(B_1\) to \(B_3\) have the optimal values reported in Table 345 and 6.

Table 3 Values of the coefficients \(B_1\) to \(B_3\) necessary to compute the coefficients \(A_1\) to \(A_3\) involved in the computation of \(R_{\mu }^{-}\)
Table 4 Values of the coefficients \(B_1\) to \(B_3\) necessary to compute the coefficients \(A_1\) to \(A_4\) involved in the computation of \(R_{\mu }^{+}\)
Table 5 Values of the coefficients \(B_1\) to \(B_3\) necessary to compute the coefficients \(A_1\) to \(A_3\) involved in the computation of \(R_{\xi }^{-}\)
Table 6 Values of the coefficients \(B_1\) to \(B_3\) necessary to compute the coefficients \(A_1\) to \(A_3\) involved in the computation of \(R_{\xi }^{+}\)

The surrogate model in Eq. 40, when fed with the parameters \(\mu _s\) and \(\xi _s^2/\sqrt{z_0}\) obtained from the empirical relations described from Eqs. 79a to 84, allows reaching the performances described in Fig. 11 in terms of residual errors on the flux footprint.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Krapez, JC., Ky, G.A. Semi-analytical Footprint Model Compliant with Arbitrary Atmospheric Stratification: Application to Monin–Obukhov Profiles. Boundary-Layer Meteorol 187, 743–791 (2023). https://doi.org/10.1007/s10546-023-00793-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10546-023-00793-2

Keywords

Navigation