Skip to main content
Log in

Global analysis of nonlinear second-mode development in a Mach-6 boundary layer from high-speed schlieren data

  • Research Article
  • Published:
Experiments in Fluids Aims and scope Submit manuscript

Abstract

The second-mode instability on a 7° half-angle sharp cone at Mach 6 is analyzed using high-speed calibrated schlieren imagery at a frame rate near the expected fundamental frequency. Experiments were conducted in the NASA Langley 20-Inch Mach 6 facility at unit Reynolds numbers between \(6.56 \; \times \; 10^{6}\) and \(9.71 \;\times \; 10^{6} \text {m}^{-1}\). Time-resolved pixel intensity signals throughout the boundary layer are reconstructed using spatially available data in the schlieren images to recover an effective sampling rate of over 10 MHz; these are then converted to quantitative density gradients using a thin-lens-based calibration technique. A global analysis is performed on the schlieren data to investigate the nonlinear growth of the second-mode fundamental and harmonic content. Pointwise measures of the autobicoherence are used to identify specific triadic interactions and the locations of their highest levels of quadratic phase coupling. Significant resonance interactions between the second-mode fundamental and harmonic instabilities were found along with interactions between these and the mean flow. Bispectral mode decomposition is employed to educe the flow structures associated with these interactions. A similar analysis is performed for the power spectrum, with power spectral densities computed for each pixel’s time-series and spectral proper orthogonal decomposition employed to derive the modal structure and energy of the flow at specific frequencies. Comparisons between the bispectral quantities and second-mode power show that nonlinear interactions, particularly resonance interactions, are closely correlated with spatiotemporal modulation of disturbances during the nonlinear stage of transition.

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
Fig. 16
Fig. 17
Fig. 18
Fig. 19

Similar content being viewed by others

Data availability

Available from authors upon request.

References

  • Ball F (1963) Energy transfer between external and internal gravity waves. J Fluid Mech 19(3):465–478

    Article  MathSciNet  ADS  Google Scholar 

  • Bathel B, Weisberger J, Herring G, King R, Jones S, Kennedy R, Laurence S (2020) Two-point, parallel-beam focused laser differential interferometry with a Nomarski prism. Appl Opt 59:244–252

    Article  CAS  PubMed  ADS  Google Scholar 

  • Berger K, Rufer S, Hollingsworth K, Wright S (2015) NASA Langley aerothermodynamics laboratory: hypersonic testing capabilities. In: AIAA Paper 2015-1337

  • Bountin D, Shiplyuk A, Maslov A (2008) Evolution of nonlinear processes in a hypersonic boundary layer on a sharp cone. J Fluid Mech 61:427–442

    Article  ADS  Google Scholar 

  • Brillinger DR (1965) An introduction to polyspectra. Ann Math Stat 36(5):1351–1374

    Article  MathSciNet  Google Scholar 

  • Butler C (2021) Interaction of second-mode disturbances with an incipiently separated compression-corner flow. J Fluid Mech 913:R4

    Article  CAS  Google Scholar 

  • Butler C, Laurence S (2021) Interaction of second-mode wave packets with an axisymmetric expansion corner. Exp Fluids 62:140

    Article  Google Scholar 

  • Butler C, Laurence S (2022) Transitional hypersonic flow over slender cone/flare geometries. J Fluid Mech 949:A37

    Article  MathSciNet  ADS  Google Scholar 

  • Chokani N (1999) Nonlinear spectral dynamics of hypersonic laminar boundary layer flow. Phys Fluids 11:3846–3851

    Article  MathSciNet  CAS  ADS  Google Scholar 

  • Chokani N (2005) Nonlinear evolution of Mack modes in a hypersonic boundary layer. Phys Fluids 17:014102

    Article  MathSciNet  ADS  Google Scholar 

  • Chou A, Leidy A, Bathel B, King R, Herring G (2018) Measurements of freestream fluctuations in the NASA Langley 20-inch Mach 6 tunnel. In: AIAA Paper 2018-3073

  • Craig S, Humble R, Hofferth J, Saric W (2019) Nonlinear behavior of the Mack modes in a hypersonic boundary layer. J Fluid Mech 872:74–99

    Article  MathSciNet  ADS  Google Scholar 

  • Gragston M, Siddiqui F, Schmisseur J (2021) Detection of second-mode instabilities on a flared cone in Mach 6 quiet flow with linear array focused laser differential interferometry. Exp Fluids 62:81

    Article  Google Scholar 

  • Hameed A, Parziale N, Paquin L, Butler C, Laurence S (2020) Spectral analysis of a hypersonic boundary layer on a right, circular cone. In: AIAA Paper 2020-362

  • Hargather M, Settles G (2012) A comparison of three quantitative Schlieren techniques. Opt Lasers Eng 50:8–17

    Article  Google Scholar 

  • Hollis B (1996) Real-gas flow properties for NASA Langley Research Center Aerothermodynamic Facilities Complex wind tunnels. No. NASA-CR-4755

  • Illingworth C (1950) Unsteady laminar flow of gas near an infinite flat plate. Math Proc Cambridge Philos Soc 46:603–613

    Article  MathSciNet  ADS  Google Scholar 

  • Kendall J (1975) Wind tunnel experiments relating to supersonic and hypersonic boundary-layer transition. AIAA J 13:290–299

    Article  ADS  Google Scholar 

  • Kennedy R (2019) An experimental investigation of hypersonic boundary-layer transition on sharp and blunt slender cones. Ph.D. thesis, University of Maryland, College Park

  • Kennedy R, Jewell J, Paredes P, Laurence S (2022) Characterization of instability mechanisms on sharp and blunt slender cones at Mach 6. J Fluid Mech 936:A39

    Article  MathSciNet  CAS  ADS  Google Scholar 

  • Kennedy R, Laurence S, Smith M, Marineau E (2017) Hypersonic boundary-layer transition features from high-speed schlieren images. In: AIAA Paper 2017-1683

  • Kennedy R, Laurence S, Smith M, Marineau E (2018) Investigation of the second-mode instability at Mach 14 using calibrated Schlieren. J Fluid Mech 845:R2

    Article  MathSciNet  ADS  Google Scholar 

  • Laurence S, Wagner A, Hannermann K (2014) Schlieren-based techniques for investigating instability development and transition in a hypersonic boundary layer. Exp Fluids 55:1–17

    Article  Google Scholar 

  • Laurence S, Wagner A, Hannermann K (2016) Experimental study of second-mode instability growth and breakdown in a hypersonic boundary layer using high-speed schlieren visualization. J Fluid Mech 797:471–503

    Article  CAS  ADS  Google Scholar 

  • Leidy A, King R, Choudhari M, Paredes P (2021) Hypersonic second-mode instability response to shaped roughness. In: AIAA Paper 2021-0149

  • Mack L (1975) Linear stability theory and the problem of the supersonic boundary-layer transition. AIAA J 13:278–289

    Article  ADS  Google Scholar 

  • Marineau E, Grossir G, Wagner A, Leinemann M, Radespiel R, Tanno H, Wagnild R, Casper K (2019) Analysis of second-mode amplitudes on sharp cones in hypersonic wind tunnels. J Spacecraft Rockets 56:307–318

    Article  CAS  ADS  Google Scholar 

  • Maslov A, Shiplyuk A, Bountin D, Sidorenko A (2006) Mach 6 boundary-layer stability experiments on sharp and blunted cones. J Spacecraft Rockets 43:71–76

    Article  ADS  Google Scholar 

  • Ort D, Dosch J (2019) Influence of mounting on the accuracy of piezoelectric pressure measurements for hypersonic boundary layer transition. In: AIAA Paper 2019-2292

  • Parziale N, Shepherd J, Hornung H (2013) Differential interferometric measurement of instability in a hypersonic boundary layer. AIAA J 51:750–754

    Article  ADS  Google Scholar 

  • Schmidt O (2020) Bispectral mode decomposition of nonlinear flows. Nonlinear Dyn 102:2479–2501

    Article  Google Scholar 

  • Shumway N, Laurence S (2015) Methods for identifying key features in schlieren images from hypersonic boundary-layer instability experiments. In: AIAA Paper 2015-1787

  • Stetson K, Thompson E, Donaldson J, Siler L (1983) Laminar boundary layer stability experiments on a cone at Mach 8, part I: sharp cone. In: AIAA Paper 1983-1761

  • Towne A, Schmidt O, Colonius T (2018) Spectral proper orthogonal decomposition and its relationship to dynamic mode decomposition and resolvent analysis. J Fluid Mech 847:821–867

    Article  MathSciNet  ADS  Google Scholar 

  • Unnikrishnan S, Gaitonde D (2020) Linear, nonlinear and transitional regimes of second-mode instability. J Fluid Mech 905:A25

    Article  MathSciNet  CAS  ADS  Google Scholar 

  • Weisberger J, Bathel B, Herring G, Buck G, Jones S, Cavone A (2020) Multi-point line focused laser differential interferometer for high-speed flow fluctuation measurements. Appl Opt 59:11180–11195

    Article  CAS  PubMed  ADS  Google Scholar 

  • Weisberger J, Bathel B, Herring G, King R, Chou A, Jones S (2019) Focused laser differential interferometry measurements at NASA Langley 20-inch Mach 6. In: AIAA Paper 2019-2903

  • Weisberger J, Bathel B, Lee J, Cavone A (2022) Linear array photodiode and data acquisition system development for multi-point line FLDI measurements. In: AIAA Paper 2022-1796

Download references

Acknowledgements

The authors gratefully acknowledge the AEDC White Oak personnel for graciously allowing NASA to use the cone model tested in the experiments as well as support from Stephen B. Jones of Analytical Mechanics Associates (AMA) for his assistance with the schlieren setup and the NASA Langley Research Center 20-Inch Mach 6 tunnel staff, including Grace Gleason (NASA Langley), Johnny Ellis (NASA Langley), Kevin Hollingsworth (Jacobs Technology), and Larson Stacey (Jacobs Technology) for conducting the facility operations.

Funding

The authors gratefully acknowledge the United States Air Force Office of Scientific Research (Dr. Brett Pokines) for support of this research through Grant FA9550-17-1-0085 as well as the NASA Hypersonic Technology Project for providing funding for testing.

Author information

Authors and Affiliations

Authors

Contributions

CES performed the analysis of schlieren data and wrote the manuscript. REK conducted the experiments. RAK supervised the experiments and provided all PCB sensor data. BFB and JMW provided significant support with experimental hardware and setup. SJL directed the research effort and objectives. All authors reviewed the manuscript.

Corresponding author

Correspondence to Stuart J. Laurence.

Ethics declarations

Conflict of interest

The authors have no competing interests to declare that are relevant to the content of this article.

Ethical approval

Not applicable.

Additional information

Publisher's Note

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

Appendix 1

Appendix 1

1.1 Effect of propagation speed on reconstructed signal

The application of the reconstruction technique in the present global analysis approach requires the assumption that all content has the same propagation speed. This is adequate for analysis of the second-mode fundamental and harmonics (Unnikrishnan and Gaitonde 2020) but will not be true in general, especially if other disturbance types (e.g., first-mode waves) are present in the boundary layer. In this section, we briefly analyze the effect of changing the propagation speed used in the reconstruction technique on the pixelwise power-spectral-density and autobicoherence.

As mentioned in Sect. 2.4, the reconstruction technique computes the linearly weighted average of the backward and forward spatial signals around the pixel of interest from images taken at times \(t_1\) and \(t_1 + \varDelta t\), where the width of spatial signals used in each image is equal to \(U_p\varDelta t\). The linearly weighted average of the spatial signals are then concatenated sequentially over all images, creating a temporal signal. An error in the propagation speed of the disturbance changes the width \(U_p\varDelta t\), resulting in a phase-shift between the backward and forward spatial signals used in the linearly weighted averaging. We will refer to this phase-shift between the spatial signals as \(\phi\).

Suppose first that equal averaging (i.e., constant weights of 1/2) was used on the backward and forward spatial signals of images at \(t_1\) and \(t_1 + \varDelta t\) and an error in the propagation speed of a disturbance wave exists. The computed average wave will have a reduced amplitude (with complete destructive interference at \(\phi =180^{\circ }\)) and a phase equal to the average phase of the two parent waves (representing a phase error from the true physical wave of \(\phi /2\)). Additionally, at the point of concatenation with the subsequent averaged wave, taken from times \(t_1 + \varDelta t\) and \(t_1 + 2\varDelta t\), there will be a discrete phase-shift in the reconstructed signal equal to \(\phi\), which occurs because the spatial signal at time \(t_1 + \varDelta t\) goes from leading the signal at time \(t_1\) to lagging the signal at \(t_1 + 2\varDelta t\). This error creates a “ringing" in the spectrum of the reconstructed signal, with artifactual spikes arising at \(nf_s+f\) and \(nf_s-f\), where n is an integer, \(f_s\) the original sampling rate (i.e., the camera frame rate), and f the disturbance frequency. This is the same mechanism that produces the spikes shown in Fig. 5, which originate from stationary irregularities in the images (e.g., window blemishes) and result in artifacts at \(nf_s\) because f is 0. In addition, spatial changes in the mean flow (e.g., the boundary-layer height) over the propagation distance used in the reconstruction will produce narrow-banded artifacts at \(nf_s\). A sufficiently high frame rate ensures that this effect is small, as changes in the mean flow state will become negligible. In these experiments, the boundary-layer growth over the schlieren field of view was estimated to lie between 2.25 and 2.75 pixels for all conditions. With an average propagation distance of approximately 38 pixels, and twice the distance used in each reconstruction procedure (incorporating forward and backward signals), this results in a minimal boundary-layer growth of 0.108 to 0.144 pixels over the spatial length used in reconstruction. Using instead a linearly weighted averaging of the backward and forward spatial signals (as is done in the present implementation) assures the averaged wave is exact at either end. This removes the discrete phase-shift at concatenation and the phase error, \(\phi\), is gradually covered over the distance \(U_p\varDelta t\). The artifacts remain but at a significantly decreased power.

If such errors in propagation speed exist, the reconstruction technique will therefore still correctly measure the disturbance frequency, but with amplitude and phase error. At \(\phi =180^{\circ }\), the weighted averaging prevents complete destructive interference (as would be the case with equal averaging), but the artifact power is at a maximum. Past 180\(^{\circ }\) phase error, the artifact power decreases but the disturbance begins to be aliased to \(f + f_s\), with complete aliasing (and complete artifact removal) occurring when an entire wavelength has been either added or removed from the spatial distance \(U_p\varDelta t\). This occurs when the error in propagation speed is equal to \(\pm \; f_s/\kappa\), where \(\kappa\) is the disturbance longitudinal wavenumber. For this reason, higher-wavenumber disturbances will experience greater amplitude and phase error from errors in their propagation speeds.

In Fig. 20, PSD and autobicoherence results (plotted with the same color scale) are presented using three different propagation speeds. The signal used is taken from condition Re97 at \(s = 330\) mm and a vertical height corresponding to the maximum second-mode disturbance (approximately 0.91 mm above the cone surface). Condition Re97 was chosen for this analysis due to the higher amplitudes of second-mode harmonics. The propagation speeds used in the reconstruction were \(0.9U_{p_0}\), \(U_{p_0}\), and \(1.1U_{p_0}\), where \(U_{p_0}\) is the calculated propagation speed shown in Table 2. A 10% error encompasses the majority of variability in the propagation speed of the disturbances of current interest. Adopting \(1.1U_{p_0}\) more accurately reconstructs disturbances propagating at the edge velocity, such as entropic and vortical disturbances. A propagation speed of \(0.9U_{p_0}\) encompasses first-mode waves (on similar geometries at Mach 6, the first-mode has been measured to have a longitudinal phase velocity of 89% of the edge velocity (Maslov et al. 2006)) and, to some degree, free-stream disturbances (which have been shown to possess propagation speeds in the range of 63–81% of the free-stream velocity in this facility (Chou et al. 2018)). Additionally, nonlinear interactions between the second-mode fundamental or harmonics with disturbances of other families, which propagate at different speeds, will produce content that propagates at an intermediate speed, \((f_1 + f_2)/(\kappa _1 + \kappa _2)\) (Ball 1963). Disturbances of other families that are likely to be interacting in this case (e.g., first-mode waves and low-frequency disturbances) will be of lower frequency and wavenumber than the second mode. Therefore, \((f_1 + f_2)/(\kappa _1 + \kappa _2)\) will be more heavily weighted toward the second-mode propagation speed and likely to fall into the range of 10% deviation from \(U_{p_0}\).

In either case of error (\(\pm \;10\%\)), there are minimal differences compared to the original results. The PSD (Fig. 20a) in the range of the fundamental second mode remains nearly unchanged in amplitude and frequency. However, the error in the PSD gets progressively worse at higher second-mode harmonics. This is expected, as the same error in propagation speed will produce larger phase errors during the signal reconstruction for higher-wavenumber disturbances. For this same reason, errors in the propagation speed also have the largest effect on higher-harmonic-producing nonlinear interactions observed in the autobicoherence. The quantitative error remains small for interactions involving the second-mode fundamental and the lower harmonics. For example, the \(f_0 + f_0 \rightarrow 2f_0\) interaction has peak \(b^2\) values of 0.84, 0.81, and 0.81 in Figs. 20b through 20d, respectively, and the \(2f_0 + f_0 \rightarrow 3f_0\) interaction has peak \(b^2\) values of 0.60, 0.57, and 0.53. The quantitative error increases for triadic interactions involving higher harmonics; however, the lobes of all notable interactions remain qualitatively consistent.

Finally, one may be concerned that, because the second-mode fundamental consists of an approximately 150 kHz band centered near the camera frame rate (286 kHz), the harmonics may instead themselves only be artifacts occurring at frequencies \(nf_s+f\) and \(nf_s-f\), where f is a fundamental frequency. Performing a quick spatial Fourier transform of the rows of pixels in the raw images discredits this, as the resulting spectra (not shown) clearly show harmonic content at the expected wavenumbers.

Fig. 20
figure 20

a PSDs using prorogation speeds of \(U_{p_0}\) (blue), \(0.9U_{p_0}\) (red), and \(1.1U_{p_0}\) (yellow). bd Autobicoherence using propagation speeds of \(U_{p_0}\) (b), \(0.9U_{p_0}\) (c), and \(1.1U_{p_0}\) (d)

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

Sousa, C.E., Kennedy, R.E., King, R.A. et al. Global analysis of nonlinear second-mode development in a Mach-6 boundary layer from high-speed schlieren data. Exp Fluids 65, 19 (2024). https://doi.org/10.1007/s00348-023-03758-w

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00348-023-03758-w

Navigation