3,638
Views
47
CrossRef citations to date
0
Altmetric
Technical Note

Estimating drag coefficient for arrays of rigid cylinders representing emergent vegetation

ORCID Icon, ORCID Icon & ORCID Icon
Pages 591-597 | Received 15 Jun 2017, Accepted 19 Jun 2018, Published online: 28 Sep 2018

ABSTRACT

Flow resistance due to vegetation is of interest for a wide variety of hydraulic engineering applications. This note evaluates several practical engineering functions for estimating bulk drag coefficient (CD) for arrays of rigid cylinders, which are commonly used to represent emergent vegetation. Many of the evaluated functions are based on an Ergun-derived expression that relates CD to two coefficients, describing viscous and inertial effects. A re-parametrization of the Ergun coefficients based on cylinder diameter (d) and solid volume fraction (φ) is presented. Estimates of CD are compared to a range of experimental data from previous studies. All functions reasonably estimate CD at low φ and high cylinder Reynolds numbers (Rd). At higher φ they typically underestimate CD. Estimates of CD utilizing the re-parametrization presented here match the experimental data better than estimates of CD made using the other functions evaluated, particularly at low φ and low Rd.

1 Introduction

Vegetation occurs in many natural and engineered water systems (O'Hare, Citation2015). In rivers the additional drag caused by vegetation acts to increase flow depths, potentially increasing the risk of flooding (Darby, Citation1999). In stormwater ponds, the resistance of vegetation has a dominant impact on the flow field and therefore affects treatment potential (Sonnenwald, Guymer, & Stovin, Citation2017). Determining vegetation drag is therefore of interest for a range of hydraulic engineering applications.

1.1 Existing measurements of CD

Arrays of rigid cylinders are often used to represent emergent vegetation, e.g. Bennett, Pirim, and Barkdoll (Citation2002), Nepf (Citation1999), Rameshwaran and Shiono (Citation2007), Rowiński and Kubrak (Citation2002), Serra, Fernando, and Rodríguez (Citation2004), Tanino and Nepf (Citation2008b) and Tinoco and Cowen (Citation2013). Table  presents seven datasets where the bulk drag coefficient, CD, for emergent cylinder arrays has been experimentally or numerically derived. The experimental and numerical methods used for determining CD are described below.

Table 1. Summary of experimental data describing drag in arrays of emergent cylinders

Traditionally, CD is obtained from experimental results for emergent cylinder arrays by equating driving forces with resistance caused by cylinders (Ferreira, Ricardo, & Franca, Citation2009; Kim & Stoesser, Citation2011; Tanino & Nepf, Citation2008a). Assuming wall and bed stresses are negligible, for emergent cylinders this equates to the balance of gravity and drag forces: (1) ρgS(1φ)=12CDaρUp2(1) where ρ is density, g is acceleration due to gravity, S is channel or energy slope, φ is solid volume fraction, a is frontal facing area (the cylinder area perpendicular to the direction of flow per unit volume, m2 m−3), and Up is mean interstitial velocity (Stone & Shen, Citation2002; Tanino, Citation2012). For cylinders φ=adπ/4 where d is cylinder diameter. In low velocities or low cylinder densities Eq. (Equation1) is impractical to apply, as it becomes difficult to measure surface slope. Bed and free surface stresses also become more important, eventually invalidating Eq. (Equation1) (Tanino & Nepf, Citation2008a). Instead, drag may be measured directly using a force sensor (Dittrich, Aberle, Schoneboom, Rodi, & Uhlmann, Citation2012; James, Goldbeck, Patini, & Jordanova, Citation2008; Tinoco & Cowen, Citation2013). Measured force is then equated directly with the the right hand side of Eq. (Equation1).

As an alternative to direct measurement, Nepf (Citation1999) assumed that turbulence production in vegetation (arrays of cylinders) is equal to dissipation, that drag dominates energy dissipation, and therefore that turbulence intensity can be equated to drag force as: (2) kUp=γ1(1φ)CDad1/3(2) where k is turbulent kinetic energy and γ1 (Tanino & Nepf, Citation2008b). Thus, instantaneous velocity measurements, e.g. from acoustic Doppler velocimetry, may be used to determine k and hence CD (Meftah & Mossa, Citation2013).

For simple geometries, such as a single cylinder or periodic arrays of cylinders, CD may be evaluated using computational fluid dynamics (CFD) tools (Kim & Stoesser, Citation2011; Koch & Ladd, Citation1997; Marjoribanks, Hardy, Lane, & Parsons, Citation2014; Rahman, Karim, & Alim, Citation2007; Stoesser, Kim, & Diplas, Citation2010), either by determining S in Eq. (Equation1) from the streamwise pressure gradient or by extracting the force on a cylinder by integrating the pressure acting on the cylinder wall.

1.2 CD estimation functions

When no physical measurements are available and CFD-based approaches are infeasible (e.g. a complex geometry) CD must be estimated. It is well established that CD for a single-cylinder is dependent on cylinder Reynolds number Rd, where Rd=Updν1 and ν is kinematic viscosity (Schlichting, Gersten, Krause, Oertel, & Mayes, Citation1960; White, Citation1991). For cylinder arrays, CD is also dependent on array characteristics (Nepf, Citation1999). Table  lists several functions that estimate CD depending on array (or vegetation) characteristics. These functions all have a basis in experimental observations and it is of interest to evaluate how successfully they estimate CD. The White (Citation1991) function is included as a base comparison.

Table 2. Equations of functions that estimate CD for arrays of emergent cylinders

The Tanino and Nepf (Citation2008a) and Tinoco and Cowen (Citation2013) functions share a common derivation. Koch and Ladd (Citation1997) showed the Ergun (Citation1952) expression for pressure drop in packed columns to successfully predict drag force. Tanino and Nepf (Citation2008a) related this expression to drag coefficient giving: (3) CD=2α0Rd+α1(3) where α0 and α1 are coefficients describing viscous and inertial drag effects respectively. Tanino and Nepf (Citation2008a) and Tinoco and Cowen (Citation2013) used their experimental CD data to estimate α0 and α1. Linking their values of α0 and α1 to the physical characteristics of their cylinder arrays led both to propose linear relationships for predicting α1 as a function of φ. Tanino and Nepf (Citation2008a) noted that α0 appeared to be independent of cylinder array characteristics and omitted the viscous term from Eq. (Equation3) in their function estimating CD. Tinoco and Cowen (Citation2013) also excluded the viscous component from their function estimating CD and suggest it is most suitable at Rd>1000. Therefore, in both functions CD is solely a function of φ.

The similarity of the methods used in the Tanino and Nepf (Citation2008a) and Tinoco and Cowen (Citation2013) studies presents an opportunity to combine their results and create enhanced estimates of CD from Eq. (Equation3). Sonnenwald, Hart, West, Stovin, and Guymer (Citation2017) re-parametrized α0 and α1 in terms of φ and d. The objectives of this note are (i) to improve the re-parametrizations of Sonnenwald et al. (Citation2017) by including additional experimental data; (ii) to demonstrate the validity of these re-parametrizations by comparing estimates of CD made using Eq. (Equation3) to experimental data; and (iii) to compare alternative estimates of CD with the re-parametrized Eq. (Equation3) and with experimental data.

2 A re-parametrization of the Ergun (Citation1952) coefficients

Figure  provides a comparison between values of α0 and α1 and the corresponding values of φ and d for a range of data. Results from Koch and Ladd (Citation1997) are plotted taking lattice units as mm for comparison purposes.

Figure 1. Plots of the Ergun (Citation1952) coefficients with respect to cylinder array characteristics, (a) α0 vs φ, (b) α0 vs d, (c) α1 vs φ, (d) α1 vs d; Meftah and Mossa (Citation2013),

Stoesser et al. (Citation2010),
Tanino and Nepf (Citation2008a) ○ Tinoco and Cowen (Citation2013), − best-fit Eq. (Equation4a), ⋄ Koch and Ladd (Citation1997) taking lattice units as mm

Figure 1. Plots of the Ergun (Citation1952) coefficients with respect to cylinder array characteristics, (a) α0 vs φ, (b) α0 vs d, (c) α1 vs φ, (d) α1 vs d; ▹ Meftah and Mossa (Citation2013), Display full size Stoesser et al. (Citation2010), Display full size Tanino and Nepf (Citation2008a) ○ Tinoco and Cowen (Citation2013), − best-fit Eq. (Equation4a(4a) α0=6475d+32(4a) ), ⋄ Koch and Ladd (Citation1997) taking lattice units as mm

Figure a does not suggest any systematic relationship between α0 and φ, which is consistent with the conclusions of Tanino and Nepf (Citation2008a).

Figure b shows a positive correlation between α0 and d. This is mainly due to the results of Tinoco and Cowen (Citation2013), who varied both φ and d. Tanino and Nepf (Citation2008a), who varied only φ, did not find a relationship with α0. It is therefore reasonable to conclude that the variation in α0 observed by Tinoco and Cowen (Citation2013) is due to d. The results of Koch and Ladd (Citation1997) show a similar trend. Together they suggest a linear relationship between α0 and d and as a result the data (excluding that of Koch & Ladd, Citation1997) presented in Fig. b have been fit to a linear function, Eq. (Equation4a), shown in Fig. b.

Tanino (Citation2012) suggested that viscous drag, the component described by α0, is proportional to d/s, where s is cylinder spacing. A linear relationship with d is consistent with this. No relationship between α0 and s (either on its own or with d) was found.

Figure c shows a positive correlation between α1 and φ, which is consistent with both Tanino and Nepf (Citation2008a) and Tinoco and Cowen (Citation2013) who both suggested a linear relationship between α1 and φ. Tanino (Citation2012) suggested that the inertial drag (described by α1) is strongly linked to flow-field heterogeneity, and that φ provides a reasonable estimate of this. Figure d also shows a positive correlation between α1 and d. A linear relationship between α1 and d is suggested by the results of Tinoco and Cowen (Citation2013) and Koch and Ladd (Citation1997), similar to α0. If α1 also depends on d, then d may serve to indicate flow-field heterogeneity.

Together, Figs c and d suggest that α1 is a function of both φ and d and all data shown in these two figures (excluding that of Koch & Ladd, Citation1997) have been used to fit a single function (not shown in Fig. ). Combining these two parameters gives a variation in values of α1 for the same d. Least-squares curve-fitting was undertaken assuming α0=f(d) and α1=f(d,φ) are linear functions giving: (4a) α0=6475d+32(4a) (4b) α1=17d+3.2φ+0.50(4b) where Eq. (Equation4a) provides an estimate of the viscous effects and Eq. (Equation4b) provides an estimate of the inertial effects of drag when used in Eq. (Equation3). Note that the coefficients to the d terms must have units m−1 to retain non-dimensionality. Root mean square error (RMSE) values of 38.0 and 0.131 were obtained respectively for α0 and α1. Substituting Eq. (Equation4) into Eq. (Equation3) gives a new function for estimating CD: (5) CD=26475d+32Rd+17d+3.2φ+0.50(5)

3 A comparison of estimates of CD against experimental results

Figure  shows estimates of CD from the functions in Table  and Eq. (Equation5) plotted for a range of Rd, a representative selection of φ and d, and with the experimental data from Table . Each sub-figure shows increasing φ. Most functions show the expected dependency of CD on Rd except the Tanino and Nepf (Citation2008a) and Tinoco and Cowen (Citation2013) functions, which exclude a viscous term and are therefore poor estimators of CD at low Rd. At φ0.02, Fig. a–d, most of the functions provide good estimates of CD for Rd>200, also suggesting that the standard CD1 is not unreasonable in this range. At φ0.01 and Rd<200, Fig.  a and b, the White (Citation1991), Ghisalberti and Nepf (Citation2004), and Cheng (Citation2012) functions produce similar underestimates of CD. All three are based on single-cylinder formulations of drag. Figure a–d show that only Eq. (Equation5) estimates CD well at Rd<200.

Figure 2. Comparison of experimental values of CD (data) to estimates (functions) at a selection of different values of φ and d (shown in top right corner of plot)

Figure 2. Comparison of experimental values of CD (data) to estimates (functions) at a selection of different values of φ and d (shown in top right corner of plot)

As φ increases, φ0.04 in Fig. e–h, the differences between the CD values estimated by each function become greater. The White (Citation1991) and Ghisalberti and Nepf (Citation2004) functions consistently underestimate CD. The Ghisalberti and Nepf (Citation2004) function predicts decreasing CD with increasing φ, which is unique among the functions presented here. The Cheng (Citation2012) function, in contrast, fits the data reasonably well at higher φ.

The differences between the Tanino and Nepf (Citation2008a) and Tinoco and Cowen (Citation2013) functions become more apparent at higher φ, with the latter estimating greater values of CD. Compared to the experimental results, the Tinoco and Cowen (Citation2013) function performs better at lower values of Rd (Fig. g) while the Tanino and Nepf (Citation2008a) function performs better at higher values (Fig. h). Despite their suggestion otherwise, the Tinoco and Cowen (Citation2013) function performs well at Rd<1000. The estimates of CD made with Eq. (Equation5) fit the data well at higher values of φ.

There are several instances where experimental configurations from the studies in Table  overlap such that measurements of CD were taken at the same φ but at different d. Figure a, b, and f show that across multiple Rd, CD increases with d. Equation (Equation5) reproduces this trend, justifying the dependence of Eq. (Equation5) on d. It is the only function that consistently fits the experimental data in Fig. .

Figure  provides a direct comparison between measured and estimated CD for each function. The Ghisalberti and Nepf (Citation2004) and White (Citation1991) functions (Fig. b and e) consistently underestimate CD with RMSE values of 3.09 and 2.40. The Tanino and Nepf (Citation2008a) function (Fig. c) performs better with an RMSE value of 1.66. The Tinoco and Cowen (Citation2013) function (Fig. d) appears to perform well with an RMSE value of 1.16, but shows significant scatter. Horizontal bands in Fig. c and d indicate that the same CD value is estimated at the same φ despite different values of d and Rd. The Cheng (Citation2012) function (Fig. a) also appears to estimate CD reasonably well, with an RMSE value of 1.28 as it performs less well at higher CD and φ.

Figure 3. Measured CD compared with estimated CD using the functions of (a) Cheng (Citation2012), (b) Ghisalberti and Nepf (Citation2004), (c) Tanino and Nepf (Citation2008a), (d) Tinoco and Cowen (Citation2013), (e) White (Citation1991), and (f) Equation (Equation5); – is a line of equality

Figure 3. Measured CD compared with estimated CD using the functions of (a) Cheng (Citation2012), (b) Ghisalberti and Nepf (Citation2004), (c) Tanino and Nepf (Citation2008a), (d) Tinoco and Cowen (Citation2013), (e) White (Citation1991), and (f) Equation (Equation5(5) CD=26475d+32Rd+17d+3.2φ+0.50(5) ); – is a line of equality

Equation (Equation5) (Fig. f) has the tightest clustering around the line of equality with an RMSE value of 0.52, showing that of the six functions evaluated it estimates values of CD closest to experimental measurements. Therefore, the dependence of α0 and α1 on φ and d suggested by Sonnenwald et al. (Citation2017) is reasonable. Note that these functions have only been tested over the range of 0.003φ0.4, 0.003d0.025, and 12Rd3838 and care must be taken applying them outside of this range.

4 Conclusions

A re-parametrization of the Ergun-derived coefficients α0 and α1 has been presented. This resulted in a function for estimating drag coefficient (CD), which has been compared to experimental data alongside several other functions estimating CD in arrays of rigid cylinders representing emergent vegetation. All functions perform well for low solid volume fractions (φ) and high cylinder Reynolds number (Rd), and generally the standard CD1 is not unreasonable here. As Rd decreases, only those functions that include viscous drag effects provide reasonable results. As φ increases, many of the functions underestimate CD. The function detailed in this study, which includes viscous effects, φ, and also a dependency on cylinder diameter (d), provides improved estimates of CD.

Notations

a=

frontal facing area (m2 m−3)

CD=

drag coefficient (–)

d=

cylinder diameter (m)

g=

gravity acceleration (m s−2)

k=

turbulent kinetic energy (m2 s−2)

Rd=

cylinder Reynolds number (–)

S=

channel slope (–)

s=

cylinder spacing (m)

Up=

mean interstitial velocity (m s−1)

α0=

coefficient describing viscous effects (m−1)

α1=

coefficient describing inertial effects (−)

γ=

turbulence intensity scaling coefficient (–)

ν=

kinematic viscosity (m2 s−1)

ρ=

density (kg m−3)

φ=

solid volume fraction (–)

Acknowledgments

The authors thank the anonymous reviewers for their feedback.

Additional information

Funding

This research was funded by the Engineering and Physical Sciences Research Council (EPSRC) [grants EP/K024442/1, EP/K025589/1 and via the Warwick Impact Acceleration Account EP/K503848/1].

References

  • Bennett, S. J., Pirim, T., & Barkdoll, B. D. (2002). Using simulated emergent vegetation to alter stream flow direction within a straight experimental channel. Geomorphology, 44(1), 115–126. doi: 10.1016/S0169-555X(01)00148-9
  • Cheng, N. S. (2012). Calculation of drag coefficient for arrays of emergent circular cylinders with pseudofluid model. Journal of Hydraulic Engineering, 139(6), 602–611. doi: 10.1061/(ASCE)HY.1943-7900.0000722
  • Darby, S. E. (1999). Effect of riparian vegetation on flow resistance and flood potential. Journal of Hydraulic Engineering, 125(5), 443–454. doi: 10.1061/(ASCE)0733-9429(1999)125:5(443)
  • Dittrich, A., Aberle, J., Schoneboom, T., Rodi, W., & Uhlmann, M. (2012). Drag forces and flow resistance of flexible riparian vegetation. In W. Rodi & M. Uhlmann (Eds.), Environmental fluid mechanics: Memorial colloquium on environmental fluid mechanics in honour of Professor Gerhard H. Jirka (pp. 195–215). London: CRC Press, Taylor & Francis Group.
  • Ergun, S. (1952). Fluid flow through packed columns. Chemical Engineering Progress, 48, 89–94.
  • Ferreira, R. M. L., Ricardo, A. M., & Franca, M. J. (2009). Discussion of “Laboratory investigation of mean drag in a random array of rigid, emergent cylinders” by Yukie Tanino and Heidi M. Nepf. Journal of Hydraulic Engineering, 135(8), 690–693. doi: 10.1061/(ASCE)HY.1943-7900.0000021
  • Ghisalberti, M., & Nepf, H. M. (2004). The limited growth of vegetated shear layers. Water Resources Research, 40(7), W07502. doi: 10.1029/2003WR002776
  • James, C. S., Goldbeck, U. K., Patini, A., & Jordanova, A. A. (2008). Influence of foliage on flow resistance of emergent vegetation. Journal of Hydraulic Research, 46(4), 536–542. doi: 10.3826/jhr.2008.3177
  • Kim, S. J., & Stoesser, T. (2011). Closure modeling and direct simulation of vegetation drag in flow through emergent vegetation. Water Resources Research, 47(10), W10511. doi: 10.1029/2011WR010561
  • Koch, D. L., & Ladd, A. J. C. (1997). Moderate Reynolds number flows through periodic and random arrays of aligned cylinders. Journal of Fluid Mechanics, 349, 31–66. doi: 10.1017/S002211209700671X
  • Marjoribanks, T., Hardy, R., Lane, S., & Parsons, D. (2014). Dynamic drag modeling of submerged aquatic vegetation canopy flows. In A. J. Schleiss, G. de Ceseare, M. J. Franca, & M. Pfister (Eds.), River Flow 2014 (pp. 517–524). London: CRC Press, Taylor & Francis Group.
  • Meftah, M. B., & Mossa, M. (2013). Prediction of channel flow characteristics through square arrays of emergent cylinders. Physics of Fluids, 25(4), 045102.
  • Nepf, H. M. (1999). Drag, turbulence, and diffusion in flow through emergent vegetation. Water Resources Research, 35(2), 479–489. doi: 10.1029/1998WR900069
  • O'Hare, M. T. (2015). Aquatic vegetation–a primer for hydrodynamic specialists. Journal of Hydraulic Research, 53(6), 687–698. doi: 10.1080/00221686.2015.1090493
  • Rahman, M. M., Karim, M. M., & Alim, M. A. (2007). Numerical investigation of unsteady flow past a circular cylinder using 2-D finite volume method. Journal of Naval Architecture and Marine Engineering, 4(1), 27–42.
  • Rameshwaran, P., & Shiono, K. (2007). Quasi two-dimensional model for straight overbank flows through emergent vegetation on floodplains. Journal of Hydraulic Research, 45(3), 302–315. doi: 10.1080/00221686.2007.9521765
  • Rowiński, P. M., & Kubrak, J. (2002). A mixing-length model for predicting vertical velocity distribution in flows through emergent vegetation. Hydrological Sciences Journal, 47(6), 893–904. doi: 10.1080/02626660209492998
  • Schlichting, H., Gersten, K., Krause, E., Oertel, H., & Mayes, K.. (1960). Boundary-layer theory (Vol. 7). Berlin: Springer.
  • Serra, T., Fernando, H. J. S., & Rodríguez, R. V. (2004). Effects of emergent vegetation on lateral diffusion in wetlands. Water Research, 38(1), 139–147. doi: 10.1016/j.watres.2003.09.009
  • Sonnenwald, F., Guymer, I., & Stovin, V. (2017). Computational fluid dynamics modelling of residence times in vegetated stormwater ponds. Proceedings of the Institution of Civil Engineers - Water Management, 171(2), 76–86.
  • Sonnenwald, F., Hart, J. R., West, P., Stovin, V. R., & Guymer, I. (2017). Transverse and longitudinal mixing in real emergent vegetation at low velocities. Water Resources Research, 53(1), 961–978. doi: 10.1002/2016WR019937
  • Stoesser, T., Kim, S. J., & Diplas, P. (2010). Turbulent flow through idealized emergent vegetation. Journal of Hydraulic Engineering, 136(12), 1003–1017. doi: 10.1061/(ASCE)HY.1943-7900.0000153
  • Stone, B. M., & Shen, H. T. (2002). Hydraulic resistance of flow in channels with cylindrical roughness. Journal of hydraulic engineering, 128(5), 500–506. doi: 10.1061/(ASCE)0733-9429(2002)128:5(500)
  • Tanino, Y. (2012). Flow and mass transport in vegetated surface waters. In C. Gualtieri & D. T. Mihailovic (Eds.), Fluid mechanics of environmental interfaces (2nd ed., pp. 369–394). Abingdon: Taylor & Francis.
  • Tanino, Y., & Nepf, H. M. (2008a). Laboratory investigation of mean drag in a random array of rigid, emergent cylinders. Journal of Hydraulic Engineering, 134(1), 34–41. doi: 10.1061/(ASCE)0733-9429(2008)134:1(34)
  • Tanino, Y., & Nepf, H. M. (2008b). Lateral dispersion in random cylinder arrays at high Reynolds number. Journal of Fluid Mechanics, 600, 339–371. doi: 10.1017/S0022112008000505
  • Tinoco, R. O., & Cowen, E. A. (2013). The direct and indirect measurement of boundary stress and drag on individual and complex arrays of elements. Experiments in Fluids, 54(4), 1–16. doi: 10.1007/s00348-013-1509-3
  • White, F. M. (1991). Viscous fluid flow. New York: McGraw-Hill.