Skip to main content
Log in

On the calculation of force from PIV data using the generalized added-mass and circulatory force decomposition

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

Abstract

To understand the forces generated on an accelerating body in a fluid flow, it is useful to have the same framework for theory and experiment. A candidate for this purpose is the decomposition of fluid-dynamic force into added-mass and circulatory components. In its generalized form applicable to viscous incompressible flows, this formulation, referred to as the GAMC formulation, is applied to planar particle image velocimetry data to calculate instantaneous forces in the present study. These estimates are compared to direct force measurements and to an alternative force formulation from impulse theory, referred to as the standard impulse formulation (SIF). The chosen test case is a nominally two-dimensional circular cylinder towed through quiescent water under three acceleration profiles with peak Reynolds numbers between 5100 and 5150. For all three motion profiles, the measured and filtered drag force is consistently greater than the calculated forces with a bias of 10–20%, but the trends are in close agreement. Inspection of the presented equations reveals that the GAMC is less sensitive to near-body vorticity data than the SIF, which has the following consequences. First, forces calculated using the GAMC formulation are less sensitive to random error in the velocity than the SIF. This benefit comes at the cost of increased sensitivity to errors in cylinder position, but the associated uncertainty is negligible in the present study. Second, the GAMC is much more tolerant to the omission of near-body vorticity data, which is an attractive feature for PIV investigations. Finally, when no data are omitted, the SIF is more sensitive to the force components induced by uncharacterized high-frequency vibrations.

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
Fig. 20
Fig. 21
Fig. 22

Similar content being viewed by others

References

  • Achenbach E (1968) Distribution of local pressure and skin friction around a circular cylinder in cross-flow up to \(\text{ Re }= 5\times 10^6\). J Fluid Mech 34(4):625–639

    Article  Google Scholar 

  • Albrecht T, del Campo V, Weier T, Metzkes H, Stiller J (2013) Deriving forces from 2D velocity field measurements. Eur Phys J Spec Top 220(1):91–100

    Article  Google Scholar 

  • DeVoria AC, Carr ZR, Ringuette MJ (2014) On calculating forces from the flow field with application to experimental volume data. J Fluid Mech 749:297–319

    Article  Google Scholar 

  • Driscoll TA, Trefethen LN (2002) Schwarz–Christoffel mapping, vol 8. Cambridge University Press, Cambridge

    Book  Google Scholar 

  • Eldredge JD (2007) Numerical simulation of the fluid dynamics of 2D rigid body motion with the vortex particle method. J Comput Phys 221(2):626–648

    Article  MathSciNet  Google Scholar 

  • Fackrell S (2011) Study of the added mass of cylinders and spheres. Ph.D. thesis, University of Windsor

  • Graham W, Pitt Ford C, Babinsky H (2017) An impulse-based approach to estimating forces in unsteady flow. J. Fluid Mech. 815:60–76

    Article  Google Scholar 

  • Kriegseis J, Rival DE (2014) Vortex force decomposition in the tip region of impulsively-started flat plates. J. Fluid Mech. 756:758–770

    Article  Google Scholar 

  • Kurtulus DF, Scarano F, David L (2007) Unsteady aerodynamic forces estimation on a square cylinder by TR-PIV. Exp. Fluids 42(2):185–196

    Article  Google Scholar 

  • Laschka B (1985) Unsteady flows—fundamentals and application. In: AGARD conference Proceedings No. 386, North Atlantic Treaty Organization

  • Limacher E, Morton C, Wood D (2018) Generalized derivation of the added-mass and circulatory forces for viscous flows. Physical Review Fluids 03:014701

    Article  Google Scholar 

  • Milne-Thompson LM (1960) Theoretical hydrodynamics, 4th edn. MacMillan, New York

    MATH  Google Scholar 

  • Moffat RJ (1988) Describing the uncertainties in experimental results. Exp Therm Fluid Sci 1:3–17

    Article  Google Scholar 

  • Mohebbian A, Rival DE (2012) Assessment of the derivative-moment transformation method for unsteady-load estimation. Exp Fluids 53(2):319–330

    Article  Google Scholar 

  • Noca F (1997) On the evaluation of time-dependent fluid-dynamic forces on bluff bodies. Ph.D. thesis, California Institute of Technology

  • Noca F, Shiels D, Jeon D (1999) A comparison of methods for evaluating time-dependent fluid dynamic forces on bodies, using only velocity fields and their derivatives. J Fluids Struct 13:551–578

    Article  Google Scholar 

  • Polet DT, Rival DE, Weymouth GD (2015) Unsteady dynamics of rapid perching manoeuvres. J Fluid Mech 767:323–341

    Article  Google Scholar 

  • Rival DE, van Oudheusden B (2017) Load-estimation techniques for unsteady incompressible flows. Exp Fluids 58(3):20

    Article  Google Scholar 

  • Saffman PG (1992) Vortex dynamics. Cambridge University Press, Cambridge

    MATH  Google Scholar 

  • Scarano F, Riethmuller ML (2000) Advances in iterative multigrid PIV image processing. Exp Fluids 29(Suppl 1):S051–S060

    Google Scholar 

  • Thom A (1929) An investigation of fluid flow in two dimensions. HM Stationery Office, Richmond

    Google Scholar 

  • Thom A (1933) The flow past circular cylinders at low speeds. Proc R Soc Lond Ser A Contain Pap Math Phys Character 141(845):651–669

    Article  Google Scholar 

  • Van Oudheusden BW, Scarano F, Casimiri EWF (2006) Non-intrusive load characterization of an airfoil using PIV. Exp Fluids 40(6):988–992. https://doi.org/10.1007/s00348-006-0149-2

    Article  Google Scholar 

  • Wang C, Eldredge JD (2013) Low-order phenomenological modeling of leading-edge vortex formation. Theor Comput Fluid Dyn 27:577–598

    Article  Google Scholar 

  • Westerweel J, Scarano F (2005) Universal outlier detection for PIV data. Exp Fluids 39(6):1096–1100

    Article  Google Scholar 

  • Wieneke B (2015) PIV uncertainty quantification from correlation statistics. Meas Sci Technol 26:074002

    Article  Google Scholar 

  • Wu JC (1981) Theory for aerodynamic force and moment in viscous flows. AIAA J 19(4):432–441

    Article  Google Scholar 

  • Wu JZ, Ma HY, Zhou MD (2015) Vortical flows. Springer, Berlin

    Book  Google Scholar 

  • Xia X, Mohseni K (2013) Lift evaluation of a two-dimensional pitching flat plate. Phys Fluids 25(9):091901

    Article  Google Scholar 

Download references

Acknowledgements

The authors thank the Natural Sciences and Engineering Research Council (NSERC) for funding this research.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Eric Limacher.

Additional information

Publisher's Note

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

Electronic supplementary material

Appendices

Appendix A: Discretization errors in numerical integration of vortical impulse

The shed-vorticity impulse, as defined in Eq. (7), is evaluated using the approximation in (11). For a single elemental area, its contribution to \({\mathbf {P}}_{\text {v}}\) is calculated as

$$\begin{aligned} {\mathbf {P}}_{{\text {v}},j} \approx h^2 \bigg [ {\mathbf {x}}_j \times {\varvec{\omega }}_j \bigg ]_V . \end{aligned}$$

This approximation is equivalent to a one term Taylor series expansion. Setting

$$\begin{aligned} {\mathbf {f}}(x,y) = {\mathbf {x}}\times {\varvec{\omega }}, \end{aligned}$$
(52)

we can express

$$\begin{aligned} {\mathbf {f}}(x,y)&= {\mathbf {f}}(x_0,y_0) + (x-x_0)\frac{\partial {\mathbf {f}}}{\partial x}\bigg |_{(x_0,y_0)} + (y-y_0)\frac{\partial {\mathbf {f}}}{\partial y}\bigg |_{(x_0,y_0)} \nonumber \\&\quad + \frac{1}{2}(x-x_0)^2 \frac{\partial ^2 {\mathbf {f}}}{\partial x^2}\bigg |_{(x_0,y_0)} +\frac{1}{2}(y-y_0)^2 \frac{\partial ^2 {\mathbf {f}}}{\partial y^2}\bigg |_{(x_0,y_0)} +\cdots \end{aligned}$$
(53)

Integrating over a square area of dimensions \(h \times h\) centred on \((x_0,y_0)\), we obtain

$$\begin{aligned} \int _{-h/2}^{h/2}\int _{-h/2}^{h/2} {\mathbf {f}}(x,y) {\text {d}}x {\text {d}}y = h^2 {\mathbf {f}}(x_0,y_0) + \frac{h^2}{24} \bigg [\frac{\partial ^2 {\mathbf {f}}}{\partial x^2} + \frac{\partial ^2 {\mathbf {f}}}{\partial y^2} \bigg ] +\cdots . \end{aligned}$$
(54)

The first-order terms in the series do not contribute to the integral, and the leading-order error, \(e_n\), introduced by truncating the series to the zeroth order term is of order

$$\begin{aligned} e_n \sim {\mathcal {O}}\bigg (\frac{1}{24}h^4 \nabla ^2({\mathbf {x}}\times {\varvec{\omega }})\bigg ). \end{aligned}$$
(55)

Appendix B: Propagating uncertainty in velocity data to uncertainty in calculated force

If some quantity M is a function of the set of variables \(\{L_j\}\), the uncertainty in M can be calculated from the uncertainties in each \(L_j\) according to (Moffat 1988):

$$\begin{aligned} \delta M = \Bigg [ \sum _j \bigg (\frac{\partial M}{\partial L_j} \bigg )^2 \delta L_j \Bigg ]^{1/2}. \end{aligned}$$
(56)

This methodology can be applied to the calculation of the vortical impulse terms, the x-components of which are calculated for each formulation using

$$\begin{aligned} P_{{\text {v}}x,{\text {SIF}}}&= h^2 \sum _j y_j \omega _j, \end{aligned}$$
(57)
$$\begin{aligned} P_{{\text {v}}x,{\text {GAMC}}}&= h^2 \sum _j \bigg (1 - \frac{D^2}{4 r_j^2}\bigg )y_j \omega _j. \end{aligned}$$
(58)

Errors in these calculated impulses result from two key sources: (1) uncertainty in the velocity vectors due to error in the PIV correlations, and (2) errors in the estimated cylinder position. Errors of the first kind can be treated as random, and the standard error propagation approach in Eq. (56) can be applied. In this approach, the error in each vorticity data point is treated as an independent random variable. In fact, neighbouring vorticity data points may have correlated errors, but the assumption that these errors are uncorrelated permits tractable analysis. By contrast, uncertainty arising due to errors in the position variable are not properly treated as uncorrelated across the domain of integration. Since the position variable is measured relative to the cylinder centre, we expect the errors in \(y_j\) to be dominated by errors in the estimated cylinder location. Thus, at any instant in time, the positional error is treated as a constant offset across the domain.

Let \(y_j\) be the measured y-position in the cylinder-fixed frame of reference and let \(Y_j\) be the true position such that

$$\begin{aligned} y_j = Y_j + e_{\text {p}}, \end{aligned}$$
(59)

where \(e_{\text {p}}\) is the error in the position estimate. The vortical impulse is then calculated by means of the SIF as

$$\begin{aligned} P_{{\text {v}}x,{\text {SIF}}} = h^2 \sum _j Y_j \omega _j + e_{\text {p}} h^2 \sum _j \omega _j, \end{aligned}$$
(60)

where h is the grid spacing in both directions. Vorticity is calculated from velocity using a central differencing scheme:

$$\begin{aligned} \omega _{i,j} = \frac{1}{2h} ( v_{i+1,j} - v_{i-1,j} - u_{i,j+1} + u_{i,j-1} ), \end{aligned}$$
(61)

where u and v are the velocities in the x- and y-directions, i and j are the x- and y-indices of the PIV dataset. The second term on the right-hand side of Eq. (60) represents the impulse error due to cylinder position error, which we will denote as \(\delta P_{x,{\text {SIF,p}}}\). We expect the total circulation to be zero in the domain around a surging cylinder, but the sum \(\sum _j \omega _j\) may not be evaluated to be identically zero due to errors in the vorticity.

Using Eqs. (56) and (61), the uncertainty in the calculated vorticity can be expressed as

$$\begin{aligned} \delta \omega _{i,j} = \frac{1}{2h} \sqrt{ \delta v_{i+1,j}^2 - \delta v_{i-1,j}^2 - \delta u_{i,j+1}^2 + \delta u_{i,j-1}^2 }, \end{aligned}$$
(62)

where \(\delta u\) and \(\delta v\) are the velocity uncertainties. Within the DaVis 8.3 software, estimates of \(\delta u\) and \(\delta v\) based on correlation statistics are furnished by the method of Wieneke (2015) (n.b. herein, the estimates furnished by the DaVis software are multiplied by two to yield uncertainties with a \(95\%\) confidence integral, i.e. covering two standard deviations of random scatter). Their method accounts for random errors which tend to blunt the correlation peak to reduce the quality of particle displacement estimates. The resulting component of impulse uncertainty, \(\delta P_{x,{\text {SIF,vel}}}\), is determined using Eqs. (56) and (57) to yield

$$\begin{aligned} \delta P_{x,{\text {SIF,vel}}} = h^2 \left[ \sum _j y_j^2 \delta \omega _j^2 \right] ^{1/2}, \end{aligned}$$
(63)

where the sum over the index j includes every data point in the two-dimensional vorticity field, i.e. \(\delta \omega _j\) in Eq. (63) is corresponds to \(\omega _{i,j}\) in Eq. (61). Let us assume that the errors \(\delta P_{x,{\text {SIF,p}}}\) and \(\delta P_{x,{\text {SIF,vel}}}\) are additive according to

$$\begin{aligned} \delta P_{{\text {v}}x,{\text {SIF}}}^2&= \delta P_{x,{\text {SIF,p}}}^2 + \delta P_{x,{\text {SIF,vel}}}^2 \nonumber \\&= e_{\text {p}}^2 h^4 \left[ \sum _j \omega _j \right] ^2 + h^4 \left[ \sum _j y_j^2 \delta \omega _j^2 \right] . \end{aligned}$$
(64)

Equation (64) yields an estimate of uncertainty in the x-impulse values calculated by means of the SIF, accounting for both random errors in the PIV correlations and errors in the estimated position of the cylinder. Estimates of y-impulse uncertainty would follow an identical pattern. Let us now repeat a similar analysis for the total vortical impulse uncertainty for the GAMC formulation. For convenience, let us define

$$\begin{aligned} y_j^\prime = \bigg (1 - \frac{D^2}{4 r_j^2}\bigg )y_j, \end{aligned}$$
(65)

where \(r_j = |{\mathbf {x}}_j|\), and \(r_j^2 = x_j^2 + y_j^2\), such that

$$\begin{aligned} P_{{\text {v}}x,{\text {GAMC}}} = h^2 \sum _j y_j^\prime \omega _j. \end{aligned}$$
(66)

Equation (66) has the same form as (57), and the uncertainty due to random error in the PIV correlations is given by

$$\begin{aligned} \delta P_{x,{\text {GAMC,vel}}}&= h^2 \left[ \sum _j y_j^{\prime 2} \delta \omega _j^2 \right] ^{1/2} \nonumber \\&= h^2 \left[ \sum _j \left( 1 - \frac{D^2}{r_j^2}\right) ^2y_j^2 \delta \omega _j^2 \right] ^{1/2}. \end{aligned}$$
(67)

Comparing Eqs. (67) and (63), it is clear that the GAMC formulation is less sensitive to random error in the PIV correlations because

$$\begin{aligned} 0< \frac{D^2}{4 r_j^2} < 1 \end{aligned}$$
(68)

for all positions in the fluid domain outside the cylinder where \(r_j> D/2\). However, the cost paid for this reduced sensitivity to random error is an increased sensitivity to error in the cylinder position estimate. As above, this error is treated by considering an offset to all location estimates in the evaluation of Eq. (67). For reasons that will become apparent, let us consider the magnitude of the total vortical impulse vector rather than solely the x-component:

$$\begin{aligned} |{\mathbf {P}}_{{\text {v}}t,{\text {GAMC}}}|&= h^2 \sum _j \left( 1 - \frac{D^2}{4 r_j^2}\right) r_j \omega _j \nonumber \\&= h^2 \sum _j \left( r_j - \frac{D^2}{4 r_j}\right) \omega _j. \end{aligned}$$
(69)

Positional error will be considered radially; let \(r_j\) be the measured radial position, and let \(R_j\) be the true radial position. These are related by the position error \(e_{\text {p}}\) by

$$\begin{aligned} r_j = R_j + e_{\text {p}}. \end{aligned}$$
(70)

Equation (69) can be rewritten as

$$\begin{aligned} |{\mathbf {P}}_{{\text {v}}t,{\text {GAMC}}}| = h^2 \sum _j \bigg (R_j + e_{\text {p}} - \frac{D^2}{4(R_j + e_{\text {p}})}\bigg ) \omega _j. \end{aligned}$$
(71)

So long as the error \(e_{\text {p}}\) is small, the error in impulse magnitude due to radial position error \(e_{\text {p}}\) can be estimated as

$$\begin{aligned} \delta |{\mathbf {P}}_{{\text {v}}t,{\text {GAMC}}}|_p \approx \frac{\partial |{\mathbf {P}}_{{\text {v}}t,{\text {GAMC}}}|}{\partial e_{\text {p}}} e_{\text {p}}, \end{aligned}$$
(72)

where

$$\begin{aligned} \frac{\partial |{\mathbf {P}}_{{\text {v}}t,{\text {GAMC}}}|}{\partial e_{\text {p}}}&= h^2 \sum _j \bigg ( 1 + \frac{D^2}{4(R_j + e_{\text {p}})^2} \bigg ) \omega _j \end{aligned}$$
(73)
$$\begin{aligned}&= h^2 \sum _j \bigg ( 1 + \frac{D^2}{4 r_j^2} \bigg ) \omega _j. \end{aligned}$$
(74)

Conservatively, let it be assumed that this magnitude error contributes entirely to the x-direction of impulse, i.e. \(\delta P_{x,{\text {GAMC,p}}} \approx \delta |{\mathbf {P}}_{{\text {v}}t,{\text {GAMC}}}|_{\text {p}}\). Finally, we arrive at

$$\begin{aligned} \delta P_{x,{\text {GAMC,p}}} \approx e_{\text {p}} h^2 \sum _j \left( 1 + \frac{D^2}{4 r_j^2} \right) \omega _j. \end{aligned}$$
(75)

As before, we take the position error and random error to be additive according to

$$\begin{aligned} \delta P_{x,{\text {GAMC}}}^2= & {} e_{\text {p}}^2 h^4 \left[ \sum _j \left( 1 + \frac{D^2}{4 r_j^2} \right) \omega _j \right] ^2 \nonumber \\&+ h^4 \left[ \sum _j \left( 1 - \frac{D^2}{4 r_j^2}\right) ^2y_j^2 \delta \omega _j^2 \right] . \end{aligned}$$
(76)

Comparison of Eqs. (64) and (76) affirms the earlier assertion that the GAMC formulation is less sensitive to random error in the PIV data but more sensitive to cylinder position error, with both conclusions resulting from the inequalities in (68).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Limacher, E., Morton, C. & Wood, D. On the calculation of force from PIV data using the generalized added-mass and circulatory force decomposition. Exp Fluids 60, 4 (2019). https://doi.org/10.1007/s00348-018-2648-3

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00348-018-2648-3

Navigation