Skip to main content

Advertisement

Log in

A direct measure of the frequency response of hot-wire anemometers: temporal resolution issues in wall-bounded turbulence

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

Abstract

Based on the need to characterise the accuracy of hot-wire anemometry (HWA) in high Reynolds number wall-bounded turbulence, we here propose a novel direct method for testing the frequency response of various systems to very high frequency velocity fluctuations (up to 50 kHz). A fully developed turbulent pipe flow is exploited to provide the input velocity perturbations. Utilising the unique capabilities of the Princeton Superpipe, it is possible to explore a variety of turbulent pipe flows at matched Reynolds numbers, but with turbulent energy in different frequency ranges. Assuming Reynolds number similarity, any differences between the appropriately scaled energy spectra for these flows should be indicative of measurement error. Having established the accuracy of this testing procedure, the response of several anemometer and probe combinations is tested. While these tests do not provide a direct or definitive comparison between different anemometers (owing to non-optimal tuning in each case), they do provide useful examples of potential frequency responses that could be encountered in HWA experiments. These results are subsequently used to predict error arising from HWA response for measurements in wall-bounded turbulent flows. For current technology, based on the results obtained here, the frequency response of under- or over-damped HWA systems can only be considered approximately flat up to 5–7 kHz. For flows with substantial turbulent energy in frequencies above this range, errors in measured turbulence quantities due to temporal resolution are increasingly likely.

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

Similar content being viewed by others

Notes

  1. It should be noted that the precise value of \(S\) is not crucial to these experiments. An estimate is good enough. Any error in \(S\) will only slightly alter the overall Reynolds number at which we conducted the experiments (along with the \(z^{+}\) and \(l^{+}\) values). More importantly, the estimate of \(S\) in no way influences how well matched the experiments are in terms of \(Re_\tau \), \(z^{+}\), \(l^{+}\) , etc.

  2. The square-wave response is measured in situ, with the probe at the centreline of the pipe with the centreline velocity \(U_{0}\) matched to the measured mean at \(z^{+} = 79\) for experiment \(e5\). Since the pipe has a turbulent core, the measured square-wave response is extracted from the turbulent signal using a triggered/conditional averaging technique—see Appendix 2.

  3. These average response curves are the mean of the \(\chi_e\) verses \(f_e\) profiles determined for experiments \(e = 1\rightarrow 4\).

References

  • Ashok A, Bailey S, Hultmark M, Smits A (2012) Hot-wire spatial resolution effects in measurements of grid-generated turbulence. Exp Fluids 53(6):1713–1722. doi:10.1007/s00348-012-1382-5

    Article  Google Scholar 

  • Bailey SC, Kunkel GJ, Hultmark M, Vallikivi M, Hill JP, Meyer KA, Tsay C, Arnold CB, Smits AJ (2010) Turbulence measurements using a nanoscale thermal anemometry probe. J Fluid Mech 663:160–179

    Article  MATH  Google Scholar 

  • Bonnet JP, de Roquefort TA (1980) Determination and optimization of frequency response of constant temperature hot-wire anemometers in supersonic flows. Rev Sci Instrum 51:234–239

    Article  Google Scholar 

  • Bruun HH (1995) Hot-wire anemometry. Oxford University Press, Oxford

    Google Scholar 

  • Chin CC, Hutchins N, Ooi ASH, Marusic I (2009) Use of direct numerical simulation (DNS) data to investigate spatial resolution issues in measurements of wall-bounded turbulence. Meas Sci Technol 20(11):115,401

    Article  Google Scholar 

  • Chin CC, Hutchins N, Ooi A, Marusic I (2011) Spatial resolution correction for hot-wire anemometry in wall turbulence. Exp Fluids 50:1443–1453

    Article  Google Scholar 

  • Comte-Bellot G (1976) Hot-wire anemometry. Annu Rev Fluid Mech 8:209–231

    Article  Google Scholar 

  • Fernholz HH, Krausse E, Nockermann M, Schober M (1995) Comparative measurements in the canonical boundary layer at \({Re}_{\delta_{2}} \leqslant 6 \times 10^{4}\) on the wall of the German–Dutch windtunnel. Phys Fluids 7(6):127

    Article  Google Scholar 

  • Freymuth P (1967) Feedback control theory for constant-temperature hot-wire anemometers. Rev Sci Instrum 38(5):677–681

    Article  Google Scholar 

  • Freymuth P (1977) Frequency response and electronic testing for constant-temperature hot-wire anemometers. J Phys E Sci Instrum 10(7):705–710

    Article  Google Scholar 

  • Hites MH (1997) Scaling of high-Reynolds number turbulent boundary layers in the National Diagnostic Facility. PhD thesis, Illinois Institute of Technology, Chicago

  • Hutchins N, Nickels TB, Marusic I, Chong MS (2009) Hot-wire spatial resolution issues in wall-bounded turbulence. J Fluid Mech 635:103–136

    Article  MATH  Google Scholar 

  • Khoo BC, Chew YT, Li GL (1995) A new method by which to determine the dynamic response of marginally elevated hot-wire anemometer probes for near-wall velocity and shear stress measurements. Meas Sci Technol 6:1399–1406

    Article  Google Scholar 

  • Kidron I (1966) Measurement of the transfer function of hot-wire and hot-film turbulence transducers. IEEE Trans Instrum Meas 15:76–81

    Article  Google Scholar 

  • Ko NWM, Ho KK (1977) On the determination of frequency response of hot-wire system. IEEE Trans Instrum Meas 26(4):360–367

    Article  Google Scholar 

  • Kunkel GJ, Arnold CB, Smits AJ (2006) Development of NSTAP: nanoscale thermal anemometry probe. AIAA-paper 2006-3718

  • Li JD, Mckeon BJ, Jiang W, Morrison JF, Smits AJ (2004) The response of hot wires in high Reynolds-number turbulent pipe flow. Meas Sci Technol 15:789–798

    Article  Google Scholar 

  • Ligrani PM, Bradshaw P (1987) Spatial resolution and measurement of turbulence in the viscous sublayer using subminiature hot-wire probes. Exp Fluids 5:407–417

    Article  Google Scholar 

  • Monkewitz PA, Duncan RD, Nagib HM (2010) Correcting hot-wire measurements of stream-wise turbulence intensity in boundary layers. Phys Fluids 22(091):701

    Google Scholar 

  • Morrison JF, McKeon BJ, Jiang W, Smits AJ (2004) Scaling of the streamwise velocity component in turbulent pipe flow. J Fluid Mech 508:99–131

    Article  MATH  Google Scholar 

  • Nickels TB, Marusic I, Hafez S, Hutchins N, Chong MS (2007) Some predictions of the attached eddy model for a high Reynolds number boundary layer. Philos Trans R Soc A 365:807–822

    Article  MATH  Google Scholar 

  • Örlü R, Alfredsson PH (2010) On spatial resolution issues related to time-averaged quantities using hot-wire anemometry. Exp Fluids 49:101–110

    Article  Google Scholar 

  • Österlund JM (1999) Experimental studies of zero pressure-gradient turbulent boundary-layer flow. PhD thesis, Department of Mechanics, Royal Institute of Technology, Stockholm

  • Oxlade AR, Valente PC, Ganapathisubramani B, Morrison JF (2012) Denoising of time-resolved piv for accurate measurement of turbulence spectra and reduced error in derivatives. Exp Fluids 53:1561–1575

    Article  Google Scholar 

  • Perry AE, Li JD (1990) Experimental support for the attached-eddy hypothesis in zero-pressure-gradient turbulent boundary layers. J Fluid Mech 218:405–438

    Article  Google Scholar 

  • Perry AE, Morrison GL (1970) Static and dynamic calibrations of constant temperature hot-wire systems. J Fluid Mech 47:765–777

    Article  Google Scholar 

  • Segalini A, Cimarelli A, Rüedi JD, Angelis ED, Talamelli A (2011a) Effect of the spatial filtering and alignment error of hot-wire probes in a wall-bounded turbulent flow. Meas Sci Technol 22(10):105,408

    Article  Google Scholar 

  • Segalini A, Örlü R, Schlatter P, Alfredsson PH, Rüedi JD, Talamelli A (2011b) A method to estimate turbulence intensity and transverse Taylor microscale in turbulent flows from spatially averaged hot-wire data. Exp Fluids 51:693–700

    Article  Google Scholar 

  • Smits AJ, Zagarola MV (2005) Applications of dense gases to model testing for aeronautical and hydrodynamic applications. Meas Sci Technol 16:1710–1715

    Article  Google Scholar 

  • Smits AJ, Monty JP, Hultmark M, Bailey SC, Hutchins N, Marusic I (2011) Spatial resolution correction for wall-bounded turbulence measurements. J Fluid Mech 676:41–53

  • Talamelli A, Persiani F, Fransson JHM, Alfredsson PH, Johansson AV, Nagib HM, Rüedi JD, Sreenivasan KR, Monkewitz PA (2009) CICLoPE—a response to the need for high reynolds number experiments. Fluid Dyn Res 41(2):021407. doi:10.1017/S0022112089002892

    Article  Google Scholar 

  • Valente PC, Vassilicos JC (2011) Comment on “Dissipation and decay of fractal-generated turbulence”. Phys Fluids 23(119):101

    Google Scholar 

  • Vallikivi M, Hultmark M, Bailey S, Smits A (2011) Turbulence measurements in pipe flow using a nano-scale thermal anemometry probe. Exp Fluids 51(6):1521–1527

    Article  Google Scholar 

  • Watmuff JH (1995) An investigation of the constant-temperature hot-wire anemometer. Exp Therm Fluid Sci 11(2):117–134

    Article  Google Scholar 

  • Weiss J, Knauss H, Wagner S (2001) Method for the determination of frequency response and signal to noise ratio for constant-temperature hot-wire anemometers. Rev Sci Instrum 72(3):1904–1909

    Article  Google Scholar 

  • Zagarola MV (1996) Mean flow scaling of turbulent pipe flow. PhD thesis, Department of Mechanical and Aerospace Engineering, Princeton University, USA

  • Zagarola MV, Smits AJ (1998) Mean-flow scaling of turbulent pipe flow. J Fluid Mech 373:33–79

    Article  MATH  Google Scholar 

Download references

Acknowledgments

N.H. and J.P.M. were supported under the Australian Research Council’s Discovery Projects (Project No. DP110102448) and Future Fellowship funding schemes (Project Nos. FT110100432, FT120100409). A.J.S was supported through ONR Grant N00014-13-1-0174.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to N. Hutchins.

Appendices

Appendix 1: Variation of response within a boundary layer

Figure 11a shows the square-wave response measured in the freestream of a boundary layer wind tunnel (the Melbourne HRNBLWT) at various mean velocities \({\overline{U}} = 10\), 3 and \(1\,\hbox {ms}^{-1}\) (black, red and blue curves, respectively). The HWA system consists of CTA1 and a standard 55P15 probe. The anemometer has not been adjusted between these experiments, and thus, any difference between the three curves in Fig. 11 is purely a result of the altered mean velocity. Figure 11b shows the corresponding normalised energy spectrum of the square-wave response, which for CTA1 is shown to give a reasonable representation of the directly determined response \(\chi \) (see Fig. 3). The variation in mean velocity over the sensor causes a pronounced change in the measured response, shifting from an under-damped system at \(10\,\hbox {ms}^{-1}\) to a clear over-damped behaviour at 1 ms\(^{-1}\). Such behaviour is to be expected from the heat transfer equation. As \({\overline{U}}\) reduces from \(10\) to \(1\,\hbox {ms}^{-1}\), the wire Reynolds number has been decreased by a factor of approximately 10 reducing the heat transfer or cooling from the wire by a factor of approximately 3. Reduced heat transfer leads to increased damping. This is evident in Fig. 11 with the system switching from under- to over-damped behaviour as the mean velocity reduces.

Fig. 11
figure 11

Change with local mean velocity of a the square-wave response and b the normalised energy spectrum of the square-wave response. Black \({\overline{U}} = 10.7\,\hbox {ms}^{-1}\), red \({\overline{U}}= 3.5\,\hbox {ms}^{-1}\), blue \({\overline{U}} = 1.0\,\hbox {ms}^{-1}\). Tuning of anemometer has not been altered (changes are due only to the change in \({\overline{U}}\))

Figure 11 suggests that it is essential to check the square-wave response across the entire expected velocity range prior to a boundary layer traverse experiment. If the frequency response is only checked at freestream conditions (say \(10\,\hbox {ms}^{-1}\)), a system could encounter substantial temporal resolution issues close to the wall as the system becomes increasingly damped. For a wall-bounded turbulent flow with a freestream velocity of \(10\,\hbox {ms}^{-1}\), local mean velocities of \(3\) and \(1\,\hbox {ms}^{-1}\) would be encountered at \(z^{+} \approx 10\) and 3. Hence, the temporal response of the system would shift from under-damped to over-damped as the probe was traversed very close to the surface.

Appendix 2: Determining the square-wave response

The square-wave responses shown throughout this investigation were obtained with the probe at the pipe centreline, with the centreline velocity and pressurisation matched to the local conditions at the probe location (which was either \(z^{+} = 29\) or 80) for the baseline experiment \(e5\). Since the pipe core is fully turbulent, the injected square-wave response is superimposed over a turbulent signal. For analysis, the signal is sampled at 1.2 MHz for a duration of 10 s. An example of this sampled signal for the under-damped CTA1 (as detailed in Figs. 12, 3) is shown in Fig. 12a, with an inset showing an enhanced detail of the square-wave response superimposed onto the turbulent core flow. To separate the square-wave response from the turbulent signal, the sampled data are filtered using a sharp spectral high-pass filter at \(f = 7\) kHz. An example of the resulting filtered signal is shown in Fig. 12b. Note that the cut-off filter was selected to remove the large-scale turbulent component from the signal. A simple threshold technique is used to extract the square-wave response. In the case shown, a threshold of 0.04 V was used detect the rising edge of the square-wave input. Regions where the rising part of the signal exceeded this threshold are shown by the dot symbols on Fig. 12b. A region of the signal surrounding this threshold crossing is then extracted and added to an ensemble average as shown in Fig. 12c. The grey curves show the square-wave responses extracted from Fig. 12b. The black curve shows the average response.

Fig. 12
figure 12

Plots illustrating the extraction of the square-wave response from a turbulent signal measured in the core. a Original time series of the velocity output from the anemometer (sampled at 1.2 MHz) showing square-wave response superimposed onto turbulent signal, inset shows an expanded detail of the response to rising step function. b Data high-pass filtered at 7 kHz, dashed line shows detection threshold and red dots show rising edge detection. c Ensemble average of detected rising edge step responses, grey lines shows the individual detections shown in plots a and b and black curve shows ensemble average over 10 s of data (approximately 10,000 individual occurrences)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hutchins, N., Monty, J.P., Hultmark, M. et al. A direct measure of the frequency response of hot-wire anemometers: temporal resolution issues in wall-bounded turbulence. Exp Fluids 56, 18 (2015). https://doi.org/10.1007/s00348-014-1856-8

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00348-014-1856-8

Keywords

Navigation