Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Extending a surface-layer C n 2 model for strongly stratified conditions utilizing a numerically generated turbulence dataset

Open Access Open Access

Abstract

In Wyngaard et al., 1971, a simple model was proposed to estimate Cn2 in the atmospheric surface layer, which only requires routine meteorological information (wind speed and temperature) as input from two heights. This Cn2 model is known to have satisfactory performance in unstable conditions; however, in stable conditions, the model only covers a relatively short range of atmospheric stabilities which significantly limits its applicability during nighttime. To mitigate this limitation, in this study we construct a new Cn2 model utilizing an extensive turbulence dataset generated by a high-fidelity numerical modeling approach (known as direct numerical simulation). The most distinguishing feature of this new Cn2 model is that it covers a wide range of atmospheric stabilities including the strongly stratified (very stable) conditions. To validate this model, approximately four weeks of Cn2 data collected at the Mauna Loa Observatory, Hawaii are used for comparison, and reasonably good agreement is found between the observed and estimated values.

© 2016 Optical Society of America

1. Introduction

Accurate estimation of optical turbulence in the lower atmosphere is essential for various ground-based optical applications (e.g., astronomical observation, laser communication and target detection) [1]. The intensity of optical turbulence is commonly quantified by the refractive index structure parameter ( Cn2, m−2/3). The Monin-Obukhov similarity theory (MOST) [2] has found widespread usage for estimating Cn2 in the surface layer (typically within tens of meters above the ground; see a brief review in [3]). Traditionally, the application of MOST requires information of turbulence statistics (e.g., momentum and sensible heat fluxes [4, 5]); however, measurements of these quantities require research-grade instruments and are not readily available. To circumvent this limitation, the gradient-based MOST approach can be utilized, which only requires routine meteorological information (wind speed and temperature) at two heights. Essentially, the gradient-based approach connects turbulence statistics and mean gradients using the so-called flux-gradient similarity relationship (e.g., the Businger – Dyer relationship [6, 7]). In light of this concept, a number of gradient-based formulations have been proposed to calculate Cn2 in the surface layer ([8–15], to name a few). For example, in Wyngaard et al. [8] (hereinafter refer to as W71), a simple Cn2 model was proposed:

CT2z4/3(θ¯z)2=gT(Rig).
Here CT2(K2m2/3) is the temperature structure parameter. z (m) is the height above the ground. θ¯(K) is the mean potential temperature. gT is the similarity function. Rig is the gradient Richardson number defined as:
Rig=gθ¯θ¯/zS2.
Here g (m s−2) is the gravitational acceleration. S=(u¯/z)2+(v¯/z)2 is the mean wind shear with u¯(m s1) and v¯(m s1) being the horizontal velocity components. Rig is a stability parameter which reflects the local balance between buoyancy and shear generation. Generally, over land, Rig is positive during nighttime (stable condition), and it becomes negative during daytime (unstable condition).

Given the gradient information (which is more easily available than turbulent fluxes) and the expression for gT, CT2 can be directly calculated using Eq. (1). For visible and near-infrared light, CT2 can be easily converted to Cn2 using:

Cn2=(7.9×105PT2)2CT2(1+0.03β)2,
where P (hPa) and T (K) are pressure and temperature, respectively, and β is the ratio of sensible and latent heat fluxes (known as the Bowen ratio) [4, 5].

Although the gradient-based model can be used to estimate Cn2 in a practical way, the most challenging part is a robust construction of similarity function gT in Eq. (1). In W71, gT was constructed using the observational data collected during a field campaign in Kansas [6, 16]:

CT2z4/3(θ¯z)2=gT(Rig)=gT(fT(ζ))={1.07(19ζ1+0.5|ζ|2/3)1/2,ζ0,(4a)0.79(0.74+4.7ζ)(1+2.5ζ3/5)1/2,ζ>0.(4b)
Here ζ is the flux-based stability parameter, which can be related to Rig through:
Rig=fT(ζ)={0.74ζ(115ζ19ζ)1/2,ζ0,(5a)ζ(0.74+4.7ζ)(1+4.7ζ)2,ζ>0.(5b)
An example of gT based on Eqs. (4) and (5) is shown in Fig. 1 where the similarity function (gT) is plotted against the stability parameter (Rig).

 figure: Fig. 1

Fig. 1 Dependence of similarity function (gT) on stability parameter (Rig). The plot is based on the similarity function proposed in W71 [8], i.e., Eqs. (4) and (5).

Download Full Size | PDF

It is well known in the literature that Cn2 estimation for stable conditions is much more challenging compared with unstable conditions (see [13, 15] for example). In the case of W71, there is a limitation that the range of application of gT is very short under stable conditions. As can be seen in Fig. 1, gT (normalized CT2) drops off sharply as Rig approaches ~0.2, implying that turbulence diminishes at this limit. However, extensive laboratory and field experiments have shown evidence that turbulence survives when Rig is larger than 0.2 (see [17, 18] for example). These findings are also corroborated by the results to be presented later in this study; we observe that Rig is larger than 0.2 in about 50% of the stably stratified conditions. It is evident that the relatively short range of gT significantly limits its applicability for estimating Cn2 during nighttime.

To mitigate the above limitation, in this study, we construct a new similarity function gT for stable conditions using an extensive dataset generated by a high-fidelity numerical modeling approach called direct numerical simulation (DNS). In the DNS approach, the Navier – Stokes (N−S) equations are solved without any averaging, filtering, or other approximations for turbulence, and the grid spacing is set to be on the order of Kolmogorov scale (centimeters or less) to capture the smallest scale of turbulent motions. Owing to its precise representation of the N−S equations, DNS is considered to be of extremely high fidelity and is utilized to provide a better physical understanding of various types of turbulent flows [19]. Recently, DNS was shown to have the potential of developing similarity functions for Cn2 given its low uncertainty compared with the observationally-based approaches [20]. More importantly, by constructing a comprehensive DNS dataset, a wide range of stabilities can be covered which provides a robust estimation of gT for Rig larger than 0.2. We would like to highlight that the novelty of this paper is that it constructs a new surface-layer Cn2 model using the high-fidelity numerical simulation approach (DNS). Moreover, we focus on the model performance in capturing Cn2 variations in strongly stratified conditions, such topic has not been addressed in the previous studies [8–15].

The structure of this paper is as follows: In Section 2, we introduce the DNS approach and summarize the computational configurations. Construction of a new similarity function gT for stable conditions using DNS is reported in Section 3. In Section 4, the validation of the newly proposed gT is conducted, followed by the summary of this paper in Section 5.

2. Direct Numerical Simulation

In this work, an extensive DNS dataset is created consisting of five simulations of stably stratified open-channel flows at bulk Reynolds number Reb = ubh/ν = 20000. Here h, ub, and ν are the channel height, the bulk (averaged) velocity in the channel, and the kinematic viscosity, respectively. This DNS dataset is an extension of our previous work [20] with a much higher Reynolds number and a wider stability range. Moreover, we utilize a newly developed DNS solver (called HERCULES [21]) for the simulations to achieve a higher numerical accuracy. The simulations are conducted by solving the normalized N−S and temperature equations:

ujxj=0,
uit+uiujxj=pxi+1Rebxjuixj+ΔPδi1+Ribθδi3,
θt+θujxj=1RebPrxjθxj.
Here ΔP is the streamwise pressure gradient driving the flow and p is the dynamic pressure. Rib=(θtopθbot)gh/ub2θtop is the bulk Richardson number with the subscripts bot and top denoting the bottom and top walls of the channel, respectively. Pr = ν/k = 0.7 is the Prandtl number with k being the thermal diffusivity. The simulation domain size is chosen to be Lx × Ly × Lz = 18h × 10h × h with 2304 × 2048 × 288 grid points in the x (streamwise), y (spanwise), and z (wall normal) directions, respectively. The other simulation configurations are similar to our previous work [20] with two exceptions. First, the constant mass flow rate condition is imposed instead of the constant pressure gradient condition. This modification allows a decreasing friction velocity during the simulations. Second, instead of fixed bottom wall temperature, all the simulations use fully developed neutrally stratified flows as initial conditions. A gradual temperature decrease is then applied at the bottom wall (by changing Rib) to mimic the cooling of the ground during nighttime. It is found that the second modification is critical for obtaining similarity function gT for Rig larger than 0.2. In this study, five normalized “cooling rates” are imposed which cover both weakly and strongly stable conditions, i.e., Rib/t = 1 × 10−3 to 5 × 10−3 with t being the non-dimensional time. Finally, the simulations are run for t = 100 and the simulation results are saved every 10 non-dimensional time. The simulations were conducted on the Army Research Laboratory High-Performance Computing (ARL HPC) system using 4096 CPU processors, and each simulation took about 80 hours.

3. Construction of New Similarity Function gT

In order to construct the dependence of gT with respect to Rig using the DNS output, we first calculate the left hand side of Eq. 1. Here CT2 can be calculated using the Corrsin’s expression [22]:

CT2=1.6ε1/3χ,
where ε and χ are the energy and temperature dissipation rate, respectively. ε and χ can be directly related to the DNS output through:
ε=νuixjuixj,χ=2kθixjθixj.
Here < · > denotes horizontal (planar) averaging, and the prime symbol represents the fluctuation of a variable with respect to its planar averaged value. Therefore, the left hand side of Eq. 1 is calculated as:
1.6νuixjuixj1/32kθixjθixjz4/3(θz)2.
Rig can be calculated using the DNS output in a similar manner.

In this paper, gT is constructed based on the DNS-generated data at different times (50 < t < 100) and height locations (0.1 < z < 0.9Hb), resulting in about 4000 samples. Here Hb is defined as the height where the local sensible heat flux drops to 10% of its surface value. Please refer to [20] for more details on ensemble calculation using the DNS output.

Figure 2 shows the dependence of similarity function (gT) on stability parameter (Rig) constructed from the DNS dataset. One can see that the DNS-based gT (black symbols) agrees reasonably well with that proposed by W71 (red dashed line) for Rig < 0.1. However, in the range Rig > 0.1, the W71 gT drops off sharply and goes to zero around Rig ≈ 0.2. In contrast, a relatively wider range of Rig is covered in the DNS-based gT. Moreover, the DNS-based gT exhibits an asymptotic behavior as Rig increases. Therefore, we fit an exponential curve based on the DNS data (50th percentile), and a new gT expression is proposed for stably stratified conditions:

CT2z4/3(θ¯z)2=gT(Rig)=0.05+1.02e14.49Rig,Rig>0.

 figure: Fig. 2

Fig. 2 Dependence of similarity function (gT) on stability parameter (Rig) for stable conditions. The symbols with error bars are the ensemble results constructed from the DNS dataset. The lower error bars, the circles, and the higher error bars represent 25th, 50th, and 75th percentile values of the ensemble results, respectively. The blue dash-dot line is the regression fit, i.e., Eq. (12), based on the 50th percentile of the DNS data (r2 = 0.981 in logarithmic coordinates). The red dashed line is the similarity function proposed in W71 [8], i.e., Eqs. (4b) and (5b).

Download Full Size | PDF

In this study, we only focus on constructing new gT for Rig > 0. For the range Rig < 0, the gT proposed in W71, i.e., Eqs. (4a) and (5a) is used in order to obtain a complete gT expression for Cn2 estimation in the following section. Note that our newly proposed gT is consistent with W71 at the neutral limit, i.e., gT = 1.07 as Rig →0.

4. Validation of the Newly Proposed gT

To validate the applicability of the newly proposed gT for a wide range of atmospheric stabilities (especially for very stable conditions), we utilize approximately four weeks of Cn2 data (July 1 to 28, 2006) collected during the MLO_CN2 campaign [23] at the Mauna Loa Observatory, Hawaii. This dataset provides 5-minute averaged time-series of wind speed, temperature, and Cn2, measured by sonic anemometers on a meteorological tower at three heights (6 m, 15 m, and 25 m above the ground). The framework for calculating Cn2 from sonic anemometers can be found in [24]. The Cn2 data at 15 m will be used for validation.

The procedure to estimate Cn2 using Eq. (12) and Eqs. (4a) and (5a) is as follows:

  1. Calculate the gradients of potential temperature and wind speed at 15 m using the second-order central-difference approach. Potential temperature can be related to temperature by θ = T(P0/P)0.286 with P0 = 1000 hPa. Since pressure was only measured at 2 m during the MLO_CN2 campaign, the hypsometric equation [25] is used to calculate P at 6 m, 15 m, and 25 m. All the data are averaged over 30 minutes. In order to avoid too small values for the wind speed and temperature gradients, thresholds are set to ensure S > 0.001 s−1 and θ¯/z>0.001K m1.
  2. Based on the gradient information, Rig is calculated using Eq. (2). CT2 is then calculated using Eq. (12) for Rig > 0. For the range Rig < 0, the gT proposed in W71 is used, i.e., Eqs. (4a) and (5a). Since analytically solving gT as a function of Rig in Eqs. (4a) and (5a) is cumbersome, we numerically calculate gT for a given Rig by interpolation.
  3. CT2 is converted to Cn2 using Eq. (3). Note that we ignore the 0.03 term in Eq. (3) following [24], since about 99% of the time the Bowen ratio (β) is found to be larger than 0.3.

Figure 3 shows the comparison of observed and estimated Cn2 at 15 m at the Mauna Loa Observatory, Hawaii. In order to illustrate the advantage of the DNS-based formulation in predicting Cn2 for a wide range of atmospheric stabilities, we use the shaded areas to denote the time periods when Rig is larger than 0.2. It is found that Rig > 0.2 in about 50% of the nighttime data. Although both W71 and DNS-based formulations predict similar Cn2 variations in Rig < 0.2 during nighttime, the former is not applicable for the strongly stratified conditions, i.e, Rig > 0.2. In the following, we will only focus on the DNS-based Cn2 model.

 figure: Fig. 3

Fig. 3 Time-series of Cn2 (15 m) at the Mauna Loa Observatory during July 1 to 28, 2006. The circles are the observational data measured at Mauna Loa Observatory, Hawaii. The black solid and hollow circles correspond to the cases Um > 2 m s−1 and Um < 2 m s−1, respectively, with Um being the magnitude of wind speed. The blue solid lines are the Cn2 values predicted using the DNS-based formulation, i.e., Eq. (12) for stable conditions and Eqs. (4a) and Eqs. (5a) for unstable conditions. The green dashed lines are the predicted Cn2 using the W71 formulation, i.e., Eqs. (4) and (5). The shaded areas denote the time periods when Rig is larger than 0.2. The mean correlation coefficients (RD and RN are for daytime and nighttime, respectively, calculated based on log10Cn2) between observation and DNS-based estimation during the four weeks are also shown in the bottom-left corner of each sub-figure. The data with Um < 2 m s−1 are excluded from the calculation of correlation coefficients. In addition, the data during July 28−29 are excluded due to the occurrence of rainfall.

Download Full Size | PDF

In terms of the performance of the DNS-based formulation, overall, the agreement between the estimated and observed Cn2 is good, especially for the first week, i.e., Fig. 3(a). As expected, the prediction of nighttime Cn2 is poorer compared with the daytime. Please refer to the correlation coefficients in Fig. 3 for each week and the overall correlation coefficients in Fig. 4(a). The most challenging part is the sharp drop-offs of Cn2 during the morning and evening transitions when θ¯/z approaches zero. Although these sharp drop-offs are qualitatively captured by the proposed approach, there are cases that their amplitudes are largely overestimated (e.g., 18 UTC on July 1 and 6). In some cases, theses sharp drop-offs are predicted earlier by about one hour (e.g., 6 UTC on July 9 and 10). In addition, in calm wind conditions, i.e., the magnitude of wind speed (Um) less than 2 m s−1, as denoted by the black hollow cycles in Fig. 3, the magnitude of observed Cn2 is found to be very oscillatory in time and hard to predict. Note that the above adverse impacts significantly degrade the correlation between observation and estimation for stable conditions, e.g., see the relatively large scatter in Fig. 4(a), especially in Cn2<1014m2/3.

 figure: Fig. 4

Fig. 4 (a) Correlation and (b) quantile-quantile plots between the observed and estimated Cn2 during July 1 to 28, 2006. RD and RN are the correlation coefficients (calculated based on log10Cn2 ) for daytime and nighttime, respectively. The calm wind data (Um < 2 m s−1) are excluded.

Download Full Size | PDF

In addition to the temporal variation of Cn2, it is important to evaluate the performance of the new Cn2 formulation in capturing the percentile distribution of the observational data. Figure 4(b) shows the quantile-quantile plot of observed and estimated Cn2 during July 1 to 28, 2006. Quite interestingly, although the correlation coefficient between observation and estimation is satisfactory during daytime, the estimated Cn2 tends to underestimate the low Cn2 values (black plus-symbols). In contrast, during nighttime, the new Cn2 formulation is found to have a relatively better performance in capturing the percentile distribution of the observed Cn2 (the blue cross-symbols).

5. Summary

In this study, we construct a new surface-layer Cn2 model utilizing a high-fidelity numerically-generated turbulence dataset. Compared with the Cn2 model proposed by Wyngaard et al. [8], the most distinguishing feature of the new model is that it covers a wide range of atmospheric stabilities including the very stable regime. To validate this newly proposed Cn2 model, we utilize approximately four weeks of observed Cn2 data collected during the MLO_CN2 campaign at Mauna Loa Observatory, Hawaii. This proposed Cn2 model is shown to have satisfactory performance in estimating Cn2 in the atmospheric surface layer during nighttime, including the strongly stratified (very stable) conditions. However, accurate prediction of temporal variation of Cn2 during morning and evening transitions is found to be very challenging, as the potential temperature gradient approaches zero. In addition, we observe that the Wyngaard’s model tends to underestimate the low Cn2 values for unstable conditions, and improvements are needed in the future work.

Acknowledgments

The authors acknowledge financial support received from the Department of Defense (AFOSR grant under award number FA9550-12-1-0449). Any opinions, findings and conclusions or recommendations expressed in this material are those of the authors and do not necessarily reflect the views of the Department of Defense. The authors also acknowledge computational resources obtained from the Department of Defense Supercomputing Resource Center (DSRC). In addition, the authors would like to thank Adam DeMarco for his valuable comments and suggestions for improving the quality of the paper, as well as his kind help on the DSRC systems.

References and links

1. L. C. Andrews and R. L. Phillips, Laser Beam Propagation through Random Media, Vol. 152 (SPIE, 2005). [CrossRef]  

2. A. S. Monin and A. M. Obukhov, “Basic laws of turbulent mixing in the surface layer of the atmosphere,” Tr. Akad. Nauk SSSR Geophiz. Inst. 24(151), 163–187 (1954).

3. R. R. Beland, Infrared and Electro-Optical Systems Handbook, Vol. 2 (SPIE, 1993).

4. R. B. Stull, An Introduction to Boundary Layer Meteorology (Kluwer Academic Publishers, 1988). [CrossRef]  

5. P. S. Arya, Introduction to Micrometeorology (Academic, 2001).

6. J. A. Businger, J. C. Wyngaard, Y. Izumi, and E. F. Bradley, “Flux-profile relationships in the atmospheric surface layer,” J. Atmos. Sci. 28, 181–189 (1971). [CrossRef]  

7. A. Dyer, “A review of flux-profile relationships,” Bound.-Layer Meteor. 7, 363–372 (1974). [CrossRef]  

8. J. C. Wyngaard, Y. Izumi, A. Stuart, and J. R. Collins, “Behavior of the refractive-index-structure parameter near the ground,” J. Opt. Soc. Am. 61, 1646–1650 (1971). [CrossRef]  

9. M. Wesely and E. Alcaraz, “Diurnal cycles of the refractive index structure function coefficient,” J. Geophys. Res. 78, 6224–6232 (1973). [CrossRef]  

10. K. Davidson, G. Schacher, C. Fairall, and A. Goroch, “Verification of the bulk method for calculating overwater optical turbulence,” Appl. Opt. 20, 2919–2924 (1981). [CrossRef]   [PubMed]  

11. K. E. Kunkel and D. L. Walters, “Modeling the diurnal dependence of the optical refractive index structure parameter,” J Geophys. Res. 88, 10,999–11,004 (1983). [CrossRef]  

12. E. L. Andreas, “Estimating Cn2 over snow and sea ice from meteorological data,” J. Opt. Soc. Am. A 5, 481–495 (1988). [CrossRef]  

13. P. A. Frederickson, K. L. Davidson, C. R. Zeisse, and C. S. Bendall, “Estimating the refractive index structure parameter (Cn2) over the ocean using bulk methods,” J. Appl. Meteor. 39, 1770–1783 (2000). [CrossRef]  

14. V. Thiermann, A. Karipot, I. Dirmhirn, P. Poschl, and C. Czekits, “Optical turbulence over paved surfaces,” Proc. SPIE 2471, 197–203 (1995). [CrossRef]  

15. A. Tunick, “Cn2 model to calculate the micrometeorological influences on the refractive index structure parameter,” Environ. Model. Softw. 18, 165–171 (2003). [CrossRef]  

16. J. Wyngaard and O. Coté, “The budgets of turbulent kinetic energy and temperature variance in the atmospheric surface layer,” J. Atmos. Sci. 28, 190–201 (1971). [CrossRef]  

17. Z. Sorbjan, “Gradient-based scales and similarity laws in the stable boundary layer,” Q. J. R. Meteorol. Soc. 136, 1243–1254 (2010).

18. A. A. Grachev, E. L. Andreas, C. W. Fairall, P. S. Guest, and P. O. G. Persson, “The critical Richardson number and limits of applicability of local similarity theory in the stable boundary layer,” Bound.-Layer Meteor. 147, 51–82 (2013). [CrossRef]  

19. P. Moin and K. Mahesh, “Direct numerical simulation: a tool in turbulence research,” Annu. Rev. Fluid Mech. 30, 539–578 (1998). [CrossRef]  

20. P. He and S. Basu, “Development of similarity relationships for energy dissipation rate and temperature structure parameter in stably stratified flows: A direct numerical simulation approach,” Environ. Fluid Mech. 16, 373–399 (2015). [CrossRef]  

21. P. He, “A high order finite difference solver for massively parallel simulations of stably stratified turbulent channel flows,” Comput. Fluids 127, 161–173 (2016). [CrossRef]  

22. S. Corrsin, “On the spectrum of isotropic temperature fluctuations in an isotropic turbulence,” J. Appl. Phys. 22, 469–473 (1951). [CrossRef]  

23. “Mauna loa seeing study. http://www.eol.ucar.edu/isf/projects/MLO_CN2/,” (2006).

24. S. Oncley and T. Horst, “Calculation of Cn2 for visible light and sound from CSAT3 sonic anemometer measurements,” avaiable at https://www.eol.ucar.edu/system/files/files/field_project/DASH04/background.pdf (2004).

25. R. B. Stull, Meteorology for Scientists and Engineers (Brooks Cole, 1999).

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (4)

Fig. 1
Fig. 1 Dependence of similarity function (gT) on stability parameter (Rig). The plot is based on the similarity function proposed in W71 [8], i.e., Eqs. (4) and (5).
Fig. 2
Fig. 2 Dependence of similarity function (gT) on stability parameter (Rig) for stable conditions. The symbols with error bars are the ensemble results constructed from the DNS dataset. The lower error bars, the circles, and the higher error bars represent 25th, 50th, and 75th percentile values of the ensemble results, respectively. The blue dash-dot line is the regression fit, i.e., Eq. (12), based on the 50th percentile of the DNS data (r2 = 0.981 in logarithmic coordinates). The red dashed line is the similarity function proposed in W71 [8], i.e., Eqs. (4b) and (5b).
Fig. 3
Fig. 3 Time-series of C n 2 (15 m) at the Mauna Loa Observatory during July 1 to 28, 2006. The circles are the observational data measured at Mauna Loa Observatory, Hawaii. The black solid and hollow circles correspond to the cases Um > 2 m s−1 and Um < 2 m s−1, respectively, with Um being the magnitude of wind speed. The blue solid lines are the C n 2 values predicted using the DNS-based formulation, i.e., Eq. (12) for stable conditions and Eqs. (4a) and Eqs. (5a) for unstable conditions. The green dashed lines are the predicted C n 2 using the W71 formulation, i.e., Eqs. (4) and (5). The shaded areas denote the time periods when Rig is larger than 0.2. The mean correlation coefficients (RD and RN are for daytime and nighttime, respectively, calculated based on log 10 C n 2) between observation and DNS-based estimation during the four weeks are also shown in the bottom-left corner of each sub-figure. The data with Um < 2 m s−1 are excluded from the calculation of correlation coefficients. In addition, the data during July 28−29 are excluded due to the occurrence of rainfall.
Fig. 4
Fig. 4 (a) Correlation and (b) quantile-quantile plots between the observed and estimated C n 2 during July 1 to 28, 2006. RD and RN are the correlation coefficients (calculated based on log 10 C n 2 ) for daytime and nighttime, respectively. The calm wind data (Um < 2 m s−1) are excluded.

Equations (12)

Equations on this page are rendered with MathJax. Learn more.

C T 2 z 4 / 3 ( θ ¯ z ) 2 = g T ( R i g ) .
R i g = g θ ¯ θ ¯ / z S 2 .
C n 2 = ( 7.9 × 10 5 P T 2 ) 2 C T 2 ( 1 + 0.03 β ) 2 ,
C T 2 z 4 / 3 ( θ ¯ z ) 2 = g T ( R i g ) = g T ( f T ( ζ ) ) = { 1.07 ( 1 9 ζ 1 + 0.5 | ζ | 2 / 3 ) 1 / 2 , ζ 0 , ( 4 a ) 0.79 ( 0.74 + 4.7 ζ ) ( 1 + 2.5 ζ 3 / 5 ) 1 / 2 , ζ > 0. ( 4 b )
R i g = f T ( ζ ) = { 0.74 ζ ( 1 15 ζ 1 9 ζ ) 1 / 2 , ζ 0 , ( 5 a ) ζ ( 0.74 + 4.7 ζ ) ( 1 + 4.7 ζ ) 2 , ζ > 0. ( 5 b )
u j x j = 0 ,
u i t + u i u j x j = p x i + 1 Re b x j u i x j + Δ P δ i 1 + R i b θ δ i 3 ,
θ t + θ u j x j = 1 Re b Pr x j θ x j .
C T 2 = 1.6 ε 1 / 3 χ ,
ε = ν u i x j u i x j , χ = 2 k θ i x j θ i x j .
1.6 ν u i x j u i x j 1 / 3 2 k θ i x j θ i x j z 4 / 3 ( θ z ) 2 .
C T 2 z 4 / 3 ( θ ¯ z ) 2 = g T ( R i g ) = 0.05 + 1.02 e 14.49 R i g , R i g > 0.
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.