Skip to main content
Log in

Representation of the Reynolds stress tensor through quadrant analysis for a near-neutral atmospheric surface layer flow

  • Original Article
  • Published:
Environmental Fluid Mechanics Aims and scope Submit manuscript

Abstract

To disseminate the role of the eddy motions on the anisotropic states of the Reynolds stress tensor (\(\varvec{B}\)), we devise a novel methodology based on the quadrant analysis, where the distributions of the invariants of \(\varvec{B}\) are studied on a \(u^{\prime }\)\(w^{\prime }\) quadrant plane, with \(u^{\prime }\) and \(w^{\prime }\) being the turbulent fluctuations in the streamwise and vertical velocities. We apply this methodology to a near-neutral atmospheric surface layer (ASL) flow, derived from a field experiment dataset having multi-level turbulence measurements. The results show that in a near-neutral ASL flow, the anisotropic states of \(\varvec{B}\) are determined by the distribution of the streamwise and cross-stream velocity variances (\(\sigma _{u}^2\) and \(\sigma _{v}^2\)) on the \(u^{\prime }\)\(w^{\prime }\) quadrant plane. By studying the contour maps of the invariants of \(\varvec{B}\) on the \(u^{\prime }\)\(w^{\prime }\) quadrant plane, we discover three distinct zones with elliptical boundaries in the \(u^{\prime }\)\(w^{\prime }\) plane, across which the anisotropic states of the eddy motions evolve. We find that the eddy motions which occur

  1. 1.

    Inside the inner elliptical zone display rod-like anisotropy being determined by \(\sigma _{v}^2\),

  2. 2.

    Within the annular zone between the inner and outer ellipses display pancake-like anisotropy being determined by both \(\sigma _{u}^2\) and \(\sigma _{v}^2\),

  3. 3.

    Outside the outer elliptical zone display rod-like anisotropy being determined by \(\sigma _{u}^2\).

We also notice that the distinction between these three zones in the \(u^{\prime }\)\(w^{\prime }\) plane is prominent at the lowest measurement level, but becomes progressively indistinguishable as the measurement height increases.

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

Similar content being viewed by others

References

  1. Pope Stephen B (2000) Turbulent flows. Cambridge University Press, Cambridge. https://doi.org/10.1017/CBO9780511840531

    Book  Google Scholar 

  2. Emory M, Iaccarino G (2014) Visualizing turbulence anisotropy in the spatial domain with componentality contours. Cent Turbul Res Ann Res Briefs 123–138

  3. Lumley John L, Newman Gary R (1977) The return to isotropy of homogeneous turbulence. J Fluid Mech 82(1):161–178. https://doi.org/10.1017/S0022112077000585

    Article  Google Scholar 

  4. Lumley John L (1979) Computational modeling of turbulent flows. Adv Appl Mech 18:123–176. https://doi.org/10.1016/S0065-2156(08)70266-7

    Article  Google Scholar 

  5. Jeffrey Alan (2001) Advanced engineering mathematics. Elsevier, Amsterdam

    Google Scholar 

  6. Choi Kwing-So, Lumley John L (2001) The return to isotropy of homogeneous turbulence. J Fluid Mech 436:59–84. https://doi.org/10.1017/S002211200100386X

    Article  Google Scholar 

  7. Nakagawa Hiroji, Nezu Iehisa (1977) Prediction of the contributions to the reynolds stress from bursting events in open-channel flows. J Fluid Mech 80(1):99–128. https://doi.org/10.1017/S0022112077001554

    Article  Google Scholar 

  8. Brugger Peter, Katul Gabriel G, De Roo Frederik, Kröniger Konstantin, Rotenberg Eyal, Rohatyn Shani, Mauder Matthias (2018) Scalewise invariant analysis of the anisotropic reynolds stress tensor for atmospheric surface layer and canopy sublayer turbulent flows. Phys Rev Fluids 3(5):054608. https://doi.org/10.1103/PhysRevFluids.3.054608

    Article  Google Scholar 

  9. Simonsen AJ, Krogstad PÅ (2005) Turbulent stress invariant analysis: clarification of existing terminology. Phys Fluids 17(8):088103. https://doi.org/10.1063/1.2009008

    Article  Google Scholar 

  10. Antonia RA, Kim J, Browne LWB (1991) Some characteristics of small-scale turbulence in a turbulent duct flow. J Fluid Mech 233:369–388. https://doi.org/10.1017/S0022112091000526

    Article  Google Scholar 

  11. Castro Ian P, Cheng Hong, Reynolds Ryan (2006) Turbulence over urban-type roughness: deductions from wind-tunnel measurements. Bound Layer Meteorol 118(1):109–131. https://doi.org/10.1007/s10546-005-5747-7

    Article  Google Scholar 

  12. Boppana VBL, Xie Z-T, Castro IP (2014) Thermal stratification effects on flow over a generic urban canopy. Bound Layer Meteorol 153(1):141–162. https://doi.org/10.1007/s10546-014-9935-1

    Article  Google Scholar 

  13. Sarkar Sutanu, Speziale Charles G (1990) A simple nonlinear model for the return to isotropy in turbulence. Phys Fluids A Fluid Dyn 2(1):84–93. https://doi.org/10.1063/1.857694

    Article  Google Scholar 

  14. Hamilton NM, Cal RB (2014) Characteristic shapes of the normalized Reynolds stress anisotropy tensor in the wakes of wind turbines with counter-rotating rotors. In: 17th international symposium on applications of laser techniques to fluid mechanics

  15. Ali Naseem, Hamilton Nicholas, Cortina Gerard, Calaf Marc, Cal Raúl Bayoán (2018) Anisotropy stress invariants of thermally stratified wind turbine array boundary layers using large eddy simulations. J Renew Sust Energy 10(1):013301. https://doi.org/10.1063/1.5016977

    Article  Google Scholar 

  16. Klipp Cheryl (2014) Turbulence anisotropy in the near-surface atmosphere and the evaluation of multiple outer length scales. Bound Layer Meteorol 151(1):57–77. https://doi.org/10.1007/s10546-013-9884-0

    Article  Google Scholar 

  17. Liu Hao, Yuan Renmin, Mei Jie, Sun Jianning, Liu Qi, Wang Yu (2017) Scale properties of anisotropic and isotropic turbulence in the urban surface layer. Bound Layer Meteorol 165(2):277–294. https://doi.org/10.1007/s10546-017-0272-z

    Article  Google Scholar 

  18. Wallace James M (2016) Quadrant analysis in turbulence research: history and evolution. Annu Rev Fluid Mech 48:131–158. https://doi.org/10.1146/annurev-fluid-122414-034550

    Article  Google Scholar 

  19. Salesky Scott T, Chamecki Marcelo, Bou-Zeid Elie (2017) On the nature of the transition between roll and cellular organization in the convective boundary layer. Bound Layer Meteorol 163(1):41–68. https://doi.org/10.1007/s10546-016-0220-3

    Article  Google Scholar 

  20. Metzger M, McKeon BJ, Holmes H (2007) The near-neutral atmospheric surface layer: turbulence and non-stationarity. Phil Trans R Soc A Math Phys Eng Sci 365(1852):859–876. https://doi.org/10.1098/rsta.2006.1946

    Article  Google Scholar 

  21. McNaughton KG, Clement RJ, Moncrieff JB (2007) Scaling properties of velocity and temperature spectra above the surface friction layer in a convective atmospheric boundary layer. Nonlinear Processes Geophys 14(3):257–271. https://doi.org/10.5194/npg-14-257-2007

    Article  Google Scholar 

  22. Vickers D, Mahrt L (2013) Quality control and flux sampling problems for tower and aircraft data. J Atmos Ocean Technol 14(3):512–526. https://doi.org/10.1175/1520-0426(1997)014<0512:QCAFSP>2.0.CO;2

    Article  Google Scholar 

  23. Wilson JD (2008) Monin-obukhov functions for standard deviations of velocity. Bound Layer Meteorol 129(3):353–369. https://doi.org/10.1007/s10546-008-9319-5

    Article  Google Scholar 

  24. Kaimal Jagadish Chandran, Finnigan John J (1994) Atmospheric boundary layer flows: their structure and measurement. Oxford University Press, Oxford

    Google Scholar 

  25. Donateo Antonio, Cava Daniela, Contini Daniele (2017) A case study of the performance of different detrending methods in turbulent-flux estimation. Bound Layer Meteorol 164(1):19–37. https://doi.org/10.1007/s10546-017-0243-4

    Article  Google Scholar 

  26. Wang Linlin, Li Dan, Gao Zhiqiu, Sun Ting, Guo Xiaofeng, Bou-Zeid Elie (2014) Turbulent transport of momentum and scalars above an urban canopy. Bound Layer Meteorol 150(3):485–511. https://doi.org/10.1007/s10546-013-9877-z

    Article  Google Scholar 

  27. Wallace James M, Eckelmann Helmut, Brodkey Robert S (1972) The wall region in turbulent shear flow. J Fluid Mech 54(1):39–48. https://doi.org/10.1017/S0022112072000515

    Article  Google Scholar 

  28. Lu SS, Willmarth WW (1973) Measurements of the structure of the reynolds stress in a turbulent boundary layer. J Fluid Mech 60(3):481–511. https://doi.org/10.1017/S0022112073000315

    Article  Google Scholar 

  29. McBean Gordon A (1974) The turbulent transfer mechanisms: a time domain analysis. Q J R Meteorol Soc 100(423):53–66. https://doi.org/10.1002/qj.49710042307

    Article  Google Scholar 

  30. Maitani T, Shaw RH (1990) Joint probability analysis of momentum and heat fluxes at a deciduous forest. Bound Layer Meteorol 52(3):283–300. https://doi.org/10.1007/BF00122091

    Article  Google Scholar 

  31. Li Dan, Bou-Zeid Elie (2011) Coherent structures and the dissimilarity of turbulent transport of momentum and scalars in the unstable atmospheric surface layer. Bound Layer Meteorol 140(2):243–262. https://doi.org/10.1007/s10546-011-9613-5

    Article  Google Scholar 

  32. Katul Gabriel, Poggi Davide, Cava Daniela, Finnigan John (2006) The relative importance of ejections and sweeps to momentum transfer in the atmospheric boundary layer. Bound Layer Meteorol 120(3):367–375. https://doi.org/10.1007/s10546-006-9064-6

    Article  Google Scholar 

  33. Nagano Y, Tagawa M (1988) Statistical characteristics of wall turbulence with a passive scalar. J Fluid Mech 196:157–185. https://doi.org/10.1017/S0022112088002654

    Article  Google Scholar 

  34. Chowdhuri S, Prabha TV (2018) An evaluation of the dissimilarity in heat and momentum transport through quadrant analysis for an unstable atmospheric surface layer flow. Environ Fluid Mech 1–30. https://doi.org/10.1007/s10652-018-9636-2

    Article  Google Scholar 

  35. Kader BA, Yaglom AM (1990) Mean fields and fluctuation moments in unstably stratified turbulent boundary layers. J Fluid Mech 212:637–662. https://doi.org/10.1017/S0022112090002129

    Article  Google Scholar 

  36. Chowdhuri, S, McNaughton, KG, Prabha TV (2018) An empirical scaling analysis of heat and momentum cospectra above the surface friction layer in a convective boundary layer. Bound Layer Meteorol 1–28. https://doi.org/10.1007/s10546-018-0397-8

    Article  Google Scholar 

  37. Drobinski P, Carlotti P, Newsom RK, Banta RM, Foster RC, Redelsperger J-L (2004) The structure of the near-neutral atmospheric surface layer. J Atmos Sci 61(6):699–714. https://doi.org/10.1175/1520-0469(2004)061<0699:TSOTNA>2.0.CO;2

    Article  Google Scholar 

  38. Bernardes M, Dias NL (2010) The alignment of the mean wind and stress vectors in the unstable surface layer. Bound Layer Meteorol 134(1):41. https://doi.org/10.1007/s10546-009-9429-8

    Article  Google Scholar 

  39. Khanna S, Brasseur JG (1998) Three-dimensional buoyancy-and shear-induced local structure of the atmospheric boundary layer. J Atmos Sci 55(5):710–743. https://doi.org/10.1175/1520-0469(1998)055<0710:TDBASI>2.0.CO;2

    Article  Google Scholar 

Download references

Acknowledgements

IITM is an autonomous institute funded by the Ministry of Earth Sciences (MoES). The authors acknowledge Dr. Keith G. McNaughton for letting them use the SLTEST data and providing the site pictures. The authors also acknowledge the three anonymous reviewers for their comments. The MATLAB codes and the SLTEST data can be made available to all the researchers by contacting the first author.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Subharthi Chowdhuri.

Additional information

Publisher's Note

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

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary material 1 (docx 700 KB)

Appendices

Appendix 1: The normalization factor while constructing \(\varvec{B}(\hat{u},\hat{w})\)

The normalization factor used in the construction of \(\varvec{B}(\hat{u}_{\mathrm{bin}}(i),\hat{w}_{\mathrm{bin}}(j))\) (Eq. 26) is the turbulent kinetic energy contained in the grid {\(\hat{u}_{\mathrm{bin}}(i)\), \(\hat{w}_{\mathrm{bin}}(j)\)}, rather than the total kinetic energy e of the half-hour run. The advantage of this normalization is it makes \(\varvec{B}(\hat{u},\hat{w})\) trace-free, and so its anisotropic states can be studied by using the two invariants only. Also since we are examining the anisotropic states of the eddy motions contained within the grid {\(\hat{u}_{\mathrm{bin}}(i)\), \(\hat{w}_{\mathrm{bin}}(j)\)}, it makes physically more sense to normalize \(\varvec{B}(\hat{u}_{\mathrm{bin}}(i),\hat{w}_{\mathrm{bin}}(j))\) by the turbulent kinetic energy contained in that grid, rather than by e. This is very similar to the spectral version of \(\varvec{B}\) (Eq. 12) as defined in Liu et al. [17], where they normalized \(\varvec{B}(\kappa )\) by the energy contained at each \(\kappa\).

One question that may arise, whether we would have obtained the original invariants of \(\varvec{B}\) (\(\xi\) and \(\eta\)) by integrating the functions \(\xi (\hat{u},\hat{w})/d\hat{u}d\hat{w}\) and \(\eta (\hat{u},\hat{w})/d\hat{u}d\hat{w}\) over the \(\hat{u}\)\(\hat{w}\) quadrant plane, if we had normalized \(\varvec{B}(\hat{u},\hat{w})\) by the total kinetic energy e. The major problem in that case would be the trace of \(\varvec{B}(\hat{u},\hat{w})\) will not be zero, thus introducing an additional invariant to study its anisotropic states. Apart from that problem, even in that case the integration of the functions \(\xi (\hat{u},\hat{w})/d\hat{u}d\hat{w}\) and \(\eta (\hat{u},\hat{w})/d\hat{u}d\hat{w}\) would not yield \(\xi\) and \(\eta\) of \(\varvec{B}\). Mathematically it can be shown that, if \(\varvec{B}(\hat{u},\hat{w})\) was normalized by e, then

$$\begin{aligned} \varvec{B}={\sum _{j=1}^{50}\sum _{i=1}^{50}} \varvec{B}(\hat{u}_{\mathrm{bin}}(i),\hat{w}_{\mathrm{bin}}(j)). \end{aligned}$$
(32)

Now combining Eqs. 2 with 8, and Eqs. 3 with 7, we can write,

$$\begin{aligned} 6\xi ^3=\mathrm {Tr}\left[ {\left( {{\sum _{j=1}^{50} \sum _{i=1}^{50}} \varvec{B}(\hat{u}_{\mathrm{bin}}(i), \hat{w}_{\mathrm{bin}}(j))}\right) }^3 \right] , \end{aligned}$$
(33)

and

$$\begin{aligned} 6\eta ^2=\mathrm {Tr}\left[ {\left( {{\sum _{j=1}^{50} \sum _{i=1}^{50}} \varvec{B}(\hat{u}_{\mathrm{bin}}(i), \hat{w}_{\mathrm{bin}}(j))}\right) }^2 \right] . \end{aligned}$$
(34)

By following the Einstein summation convention, we can expand the terms in Eqs. 33 and 34 as,

$$\begin{aligned} 6\xi ^3=\mathrm {Tr}\left[ \varvec{B}(\hat{u}_{\mathrm{bin}}(i), \hat{w}_{\mathrm{bin}}(j))\ \varvec{B} (\hat{u}_{\mathrm{bin}}(j), \hat{w}_{\mathrm{bin}}(k)) \ \varvec{B}(\hat{u}_{\mathrm{bin}}(k), \hat{w}_{\mathrm{bin}}(i))\right] , \end{aligned}$$
(35)

and

$$\begin{aligned} 6\eta ^2=\mathrm {Tr}\left[ \varvec{B}(\hat{u}_{\mathrm{bin}}(i), \hat{w}_{\mathrm{bin}}(j)) \varvec{B} (\hat{u}_{\mathrm{bin}}(j),\hat{w}_{\mathrm{bin}}(i)) \right] . \end{aligned}$$
(36)

Since the cross-terms in the Eqs. 35 and 36 do not vanish, the invariants \(\xi\) and \(\eta\) are simply not the sums of the invariants of \(\varvec{B}(\hat{u}_{\mathrm{bin}}(i),\hat{w}_{\mathrm{bin}}(j))\), and so the integration of the functions \(\xi (\hat{u},\hat{w})/d\hat{u}d\hat{w}\) and \(\eta (\hat{u},\hat{w})/d\hat{u}d\hat{w}\) do not yield the values of \(\xi\) and \(\eta\) of \(\varvec{B}\).

Appendix 2: Analytical expressions of the Reynolds stress invariants \(\xi\) and \(\eta\)

From the Cayley-Hamilton theorem [4], the characteristic polynomial of \(\varvec{B}\) can be written as,

$$\begin{aligned} {\varvec{B}}^{3}-{\mathrm {Tr}(\varvec{B}}) {\varvec{B}}^{2} +\frac{1}{2}[{({\mathrm {Tr}(\varvec{B}}))}^2 -\mathrm {Tr}({\varvec{B}}^{2})]{\varvec{B}} -\det (\varvec{B})I_{3,3}=0_{3,3}, \end{aligned}$$
(37)

where \(I_{3,3}\) and \(0_{3,3}\) are the identity and null matrices of the order \(3 \times 3\) respectively. The coefficients of the characteristic polynomial of \(\varvec{B}\) are the invariants of \(\varvec{B}\) (I, II, and III), since these quantities are conserved under any coordinate transformation. Given the fact that \(\varvec{B}\) has zero trace this makes \(I=0\), and the other two invariants of Eq. 37 (II and III) can be written as,

$$\begin{aligned} II=-\frac{1}{2} \mathrm {Tr}({\varvec{B}}^{2}), \end{aligned}$$
(38)

and

$$\begin{aligned} III=\det (\varvec{B}). \end{aligned}$$
(39)

It is to note that, since \(\varvec{B}\) is symmetric (\(\varvec{B}_{ij}=\varvec{B}_{ji}\)) Eq. 38 can also be written as,

$$\begin{aligned} II=-\frac{1}{2} (\varvec{B:B}), \end{aligned}$$
(40)

where \(\varvec{B:B}\) is the Frobenius inner product between the \(\varvec{B}\)s. Now from the matrix algebra [5], the \(\det (\varvec{B})\) can be written as,

$$\begin{aligned} \det (\varvec{B})=\frac{1}{6}[{({\mathrm {Tr}(\varvec{B}}))}^3 -3\mathrm {Tr}({\varvec{B}}^{2})\mathrm {Tr}(\varvec{B}) +2\mathrm {Tr}({\varvec{B}}^{3})]. \end{aligned}$$
(41)

Since \(\mathrm {Tr}(\varvec{B})=0\), by combining Eqs. 39 and 41 we can write,

$$\begin{aligned} III=\frac{1}{3} \mathrm {Tr}({\varvec{B}}^{3}). \end{aligned}$$
(42)

By converting the invariants II and III to \(\xi\) and \(\eta\) from Eqs. 7 and 8, we can write,

$$\begin{aligned} 6\eta ^{2} &= \left( \frac{\overline{u^{\prime 2}}}{2e} - \frac{1}{3}\right) ^{2} + \left( \frac{\overline{v^{\prime 2}}}{2e} - \frac{1}{3}\right) ^{2} + \left( \frac{\overline{w^{\prime 2}}}{2e} - \frac{1}{3}\right) ^{2} \nonumber \\&\quad +\,2\left[ \left( \frac{\overline{{u}^{\prime }{v}^{\prime }}}{2e}\right) ^{2} + \left( \frac{\overline{{v}^{\prime }{w}^{\prime }}}{2e}\right) ^{2} + \left( \frac{\overline{{u}^{\prime }{w}^{\prime }}}{2e}\right) ^{2}\right] , \end{aligned}$$
(43)

and

$$\begin{aligned} 6\xi ^{3} &= \left( \frac{\overline{u^{\prime 2}}}{2e} - \frac{1}{3}\right) ^{3} + \left( \frac{\overline{v^{\prime 2}}}{2e} - \frac{1}{3}\right) ^{3} + \left( \frac{\overline{w^{\prime 2}}}{2e} - \frac{1}{3}\right) ^{3} \nonumber \\&\quad +\,3\left( \frac{\overline{{u}^{\prime }{v}^{\prime }}}{2e}\right) ^{2} \left[ \frac{\overline{u^{\prime 2}}}{2e} + \frac{\overline{v^{\prime 2}}}{2e} - \frac{2}{3}\right] \nonumber \\&\quad +\, 3\left( \frac{\overline{{v}^{\prime }{w}^{\prime }}}{2e}\right) ^{2} \left[ \frac{\overline{v^{\prime 2}}}{2e} + \frac{\overline{w^{\prime 2}}}{2e} - \frac{2}{3}\right] \nonumber \\&\quad +\,3\left( \frac{\overline{{u}^{\prime }{w}^{\prime }}}{2e}\right) ^{2} \left[ \frac{\overline{u^{\prime 2}}}{2e} + \frac{\overline{w^{\prime 2}}}{2e} - \frac{2}{3}\right] \nonumber \\&\quad +\,6\left( \frac{\overline{u^{\prime }v^{\prime }}}{2e}\right) \left( \frac{\overline{v^{\prime }w^{\prime }}}{2e}\right) \left( \frac{\overline{u^{\prime }w^{\prime }}}{2e}\right) , \end{aligned}$$
(44)

respectively, by expanding \(\mathrm {Tr}({\varvec{B}}^{2})\) and \(\mathrm {Tr}({\varvec{B}}^{3})\) in terms of the components of \(\varvec{B}\). From Fig. 9, we notice that the components \({\overline{{u}^{\prime }{v}^{\prime }}}/{2e}\), \({\overline{{v}^{\prime }{w}^{\prime }}}/{2e}\), and \({\overline{{u}^{\prime }{w}^{\prime }}}/{2e}\) contribute negligibly to \(\varvec{B}\). We can thus simplify Eqs. 43 and 44 as,

$$\begin{aligned} 6\eta ^{2} \approx \left( \frac{\overline{u^{\prime 2}}}{2e} -\frac{1}{3}\right) ^{2} + \left( \frac{\overline{v^{\prime 2}}}{2e} -\frac{1}{3}\right) ^{2} + \left( \frac{\overline{w^{\prime 2}}}{2e} -\frac{1}{3}\right) ^{2}, \end{aligned}$$
(45)

and

$$\begin{aligned} 6\xi ^{3} \approx \left( \frac{\overline{u^{\prime 2}}}{2e} -\frac{1}{3}\right) ^{3} + \left( \frac{\overline{v^{\prime 2}}}{2e} -\frac{1}{3}\right) ^{3} + \left( \frac{\overline{w^{\prime 2}}}{2e} -\frac{1}{3}\right) ^{3}. \end{aligned}$$
(46)
Fig. 9
figure 9

The vertical profiles of each component of \(\varvec{B}\) such as, the principal stresses a\(\overline{{u^{\prime }}^2}/2e\), \(\overline{{v^{\prime }}^2}/2e\), and \(\overline{{w^{\prime }}^2}/2e\), and the cross terms b\(\overline{u^{\prime }w^{\prime }}/2e\), \(\overline{v^{\prime }w^{\prime }}/2e\), and \(\overline{u^{\prime }v^{\prime }}/2e\). The vertical profiles are averaged with the error bars showing one standard deviation from the mean

Appendix 3: Construction of a conditionally sampled \(\varvec{B}\)

In Fig. 4b, d and in Fig. 7d–f, we show the Lumley triangle where the invariants of \(\varvec{B}\) are plotted for the conditionally sampled eddy motions. We can write the elements of \(\varvec{B}\) for any conditionally sampled eddy motions as,

$$\begin{aligned} \varvec{B}_{11}(C) &= \frac{{(\sum {{u^{\prime }}^2})}_{C}}{{(\sum {{u^{\prime }}^2})}_{C}+{(\sum {{v^{\prime }}^2})}_{C} +{(\sum {{w^{\prime }}^2})}_{C}} - \frac{1}{3}\nonumber \\ \varvec{B}_{12}(C) &= \frac{{(\sum {u^{\prime }v^{\prime }})}_{C}}{{(\sum {{u^{\prime }}^2})}_{C}+{(\sum {{v^{\prime }}^2})}_{C} +{(\sum {{w^{\prime }}^2})}_{C}}\nonumber \\ \varvec{B}_{13}(C) &= \frac{{(\sum {u^{\prime }w^{\prime }})}_{C}}{{(\sum {{u^{\prime }}^2})}_{C}+{(\sum {{v^{\prime }}^2})}_{C} +{(\sum {{w^{\prime }}^2})}_{C}}\nonumber \\ \varvec{B}_{21}(C) &= \varvec{B}_{12}(C) \nonumber \\ \varvec{B}_{22}(C) &= \frac{{(\sum {{v^{\prime }}^2})}_{C}}{{(\sum {{u^{\prime }}^2})}_{C}+{(\sum {{v^{\prime }}^2})}_{C} +{(\sum {{w^{\prime }}^2})}_{C}} - \frac{1}{3}\nonumber \\ \varvec{B}_{23}(C) &= \frac{{(\sum {v^{\prime }w^{\prime }})}_{C}}{{(\sum {{u^{\prime }}^2})}_{C}+{(\sum {{v^{\prime }}^2})}_{C} +{(\sum {{w^{\prime }}^2})}_{C}} \nonumber \\ \varvec{B}_{31}(C) &= \varvec{B}_{13}(C) \nonumber \\ \varvec{B}_{32}(C) &= \varvec{B}_{23}(C) \nonumber \\ \varvec{B}_{33}(C) &= \frac{{(\sum {{w^{\prime }}^2})}_{C}}{{(\sum {{u^{\prime }}^2})}_{C}+{(\sum {{v^{\prime }}^2})}_{C} +{(\sum {{w^{\prime }}^2})}_{C}} - \frac{1}{3}, \end{aligned}$$
(47)

where C is the condition under which the eddy motions are sampled, which for example can be those motions located inside or outside the HH (Eq. 30), or within any of the three elliptical zones as defined in Eq. 31. It is to be noted here that similar to Eq. 26 we normalize the elements of \(\varvec{B}(C)\) in Eq. 47 by the turbulent kinetic energy contained within these conditionally sampled eddy motions, for the same reasons as discussed earlier.

We can write the invariants of \(\varvec{B}(C)\) as,

$$\begin{aligned} \eta (C)={\left( \frac{\lambda _{1C}^2+\lambda _{2C}^2 +\lambda _{3C}^2}{6}\right) }^{\frac{1}{2}}, \end{aligned}$$
(48)

and

$$\begin{aligned} \xi (C)={\left( \frac{\lambda _{1C}^3+\lambda _{2C}^3 +\lambda _{3C}^3}{6}\right) }^{\frac{1}{3}}, \end{aligned}$$
(49)

where \(\lambda _{1C}\), \(\lambda _{2C}\), and \(\lambda _{3C}\) are the eigenvalues of \(\varvec{B}(C)\). These invariants \(\eta (C)\) and \(\xi (C)\) are plotted on the Lumley triangle as shown in Fig. 4b, d and in Fig. 7d–f.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chowdhuri, S., Deb Burman, P.K. Representation of the Reynolds stress tensor through quadrant analysis for a near-neutral atmospheric surface layer flow. Environ Fluid Mech 20, 51–75 (2020). https://doi.org/10.1007/s10652-019-09689-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10652-019-09689-7

Keywords

Navigation