wave-frequency flows within a near-bed vegetation canopy · 2017-03-22 · (weitzman et al., 2015)...

29
Wave-frequency Flows Within a Near-bed Vegetation Canopy Henderson, Stephen M. a,, Norris, Benjamin K. b , Mullarney, Julia C. b , Bryan, Karin, R. b a School of the Environment, Washington State University (Vancouver). Vancouver, WA 98686, USA b Coastal Marine Group, Faculty of Science and Engineering, University of Waikato, Hamilton, 3240, New Zealand Abstract We study water flows and wave dissipation within near-bed pneumatophore canopies at the wave-exposed fringe of a mangrove forest on C` u Lao Dung Island, in the Mekong Delta. To evaluate canopy drag, the three-dimensional geometry of pneu- matophore stems growing upward from the buried lateral roots of Sonneratia case- olaris mangroves was reconstructed from photogrammetric surveys. In cases where hydrodynamic measurements were obtained, up to 84 stems per square meter were observed, with stem heights < 0.6 m, and basal diameters 0.01 – 0.02 m. The pa- rameter a = (frontal area of pneumatophores blocking the flow)/(canopy volume) ranged from zero to 1.8 m -1 . Within-canopy water velocity displayed a phase lead and slight attenuation relative to above-canopy flows. The phase lead was frequency-dependent, ranging from 0 – 30 degrees at the frequencies of energetic waves (> 0.1 Hz), and up to 90 degrees at lower frequencies. A model is developed for wave-induced flows within the vertically variable canopy. Scaling suggests that acceleration-induced forces and vertical mixing were negligible at wave frequen- cies. Consistent with theory, drag-induced vertical variability in velocity scaled with f = T w /(2T f ), where T w = wave period, T f =2/(C D a|u|) is the frictional time scale, C D 2 is the drag coecient, and |u| is a typical flow speed. For given wave conditions (|u| and T w ), theory predicts increasing dissipation with increasing vegetation density (i.e. increasing a), until a maximum is reached for order-one f . For larger f , within-canopy flow is so inhibited by drag that further increases in a reduce within-canopy dissipation. For observed cases, f 0.38 at energetic wave frequencies, so wave dissipation near the forest edge is expected corresponding author email: steve [email protected]; ph: USA (503) 706 8819 Preprint submitted to Continental Shelf Research October 26, 2016

Upload: others

Post on 29-Jul-2020

8 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

Wave-frequency Flows Within a Near-bed

Vegetation Canopy

Henderson, Stephen M.a,⇤, Norris, Benjamin K.b, Mullarney, Julia C.b,Bryan, Karin, R.b

aSchool of the Environment, Washington State University (Vancouver). Vancouver, WA98686, USA

bCoastal Marine Group, Faculty of Science and Engineering, University of Waikato,Hamilton, 3240, New Zealand

Abstract

We study water flows and wave dissipation within near-bed pneumatophore canopiesat the wave-exposed fringe of a mangrove forest on Cu Lao Dung Island, in theMekong Delta. To evaluate canopy drag, the three-dimensional geometry of pneu-matophore stems growing upward from the buried lateral roots of Sonneratia case-olaris mangroves was reconstructed from photogrammetric surveys. In cases wherehydrodynamic measurements were obtained, up to 84 stems per square meter wereobserved, with stem heights < 0.6m, and basal diameters 0.01 – 0.02m. The pa-rameter a =(frontal area of pneumatophores blocking the flow)/(canopy volume)ranged from zero to 1.8m�1. Within-canopy water velocity displayed a phaselead and slight attenuation relative to above-canopy flows. The phase lead wasfrequency-dependent, ranging from 0 – 30 degrees at the frequencies of energeticwaves (> 0.1Hz), and up to 90 degrees at lower frequencies. A model is developedfor wave-induced flows within the vertically variable canopy. Scaling suggests thatacceleration-induced forces and vertical mixing were negligible at wave frequen-cies. Consistent with theory, drag-induced vertical variability in velocity scaledwith ✏f = Tw/(2⇡Tf ), where Tw =wave period, Tf = 2/(CDa|u|) is the frictionaltime scale, CD ⇡ 2 is the drag coe�cient, and |u| is a typical flow speed. Forgiven wave conditions (|u| and Tw), theory predicts increasing dissipation withincreasing vegetation density (i.e. increasing a), until a maximum is reached fororder-one ✏f . For larger ✏f , within-canopy flow is so inhibited by drag that furtherincreases in a reduce within-canopy dissipation. For observed cases, ✏f 0.38 atenergetic wave frequencies, so wave dissipation near the forest edge is expected

⇤corresponding author email: steve [email protected]; ph: USA (503) 706 8819

Preprint submitted to Continental Shelf Research October 26, 2016

Page 2: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

to increase with increasing pneumatophore canopy density, but under di↵erentwave conditions the most dense canopies may occasionally approach the dissipa-tion maximum (✏f ⇡ 1). Predicted dissipation by the pneumatophore canopy wassu�cient to attenuate most wave energy over distances slightly less (more) than100m into the marsh in 1m (2m) water depth.

Keywords:wave dissipation, drag coe�cient, mangrove, photogrammetry, estuaries,vegetation

2

Page 3: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

Highlights

• Water velocity measured under waves in a Mekong Delta near-bed veg-etation canopy.

• Within-canopy velocity consistent with a simple theory incorporatingcanopy drag.

• Theory predicts dissipation proportional to canopy density for observedcases.

• At higher canopy densities, predicted wave dissipation would decline.

• Predicted wave height halved tens (hundreds) of m into swamp in 1 m(2 m) depth.

1. Introduction

Aquatic vegetation can shelter coastlines from energetic waves currents(Jadhav et al., 2013; Ri↵e et al., 2011; Temmerman et al., 2013), and some-times creates low-energy regions of sediment deposition (Bouma et al., 2007;Furukawa et al., 1997). Diverse coastal wetland ecosystems (Greenberg et al.,2006) are often productive regions of rapid carbon burial (Siikamaki et al.,2012), and many wetlands are threatened by pollution, land reclamation,and conversion to aquaculture. Mangrove swamps in particular have beenimpacted by widespread deforestation (Giri et al., 2011; Thu and Populus,2007). In Vietnam’s Mekong Delta, reduced sediment supply is expectedto compound the e↵ects of rising sea levels on coastal evolution (Anthonyet al., 2015; Nicholls et al., 1999). Improved understanding of interactinghydrodynamics and morphodynamics may prove relevant to management ofmangrove ecosystems, coastal flooding and erosion.

We present observations from a mangrove-covered coast in the SouthernMekong Delta. This coast is exposed to the open ocean, and therefore toenergetic waves. Previous researchers have measured attenuation of wavespropagating into mangrove forests, finding substantial dissipation after tensor hundreds of meters of propagation (Bao, 2011; Horstman et al., 2014;Mazda et al., 2006). A variety of models have been developed to predictwave dissipation within vegetation. Dalrymple et al. (1984) modeled thedrag forces exerted by vegetation, and the associated wave dissipation, us-ing frictionless linear wave theory to estimate the wave-induced water flows

3

Page 4: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

(hereafter ‘wave orbital velocities’) past plant stems. This approach providesa good approximation in su�ciently sparse canopies, but frictionless wavetheory is not applicable at very high canopy densities. For near-bed vegeta-tion, Lowe et al. (2005) (hereafter LKM05) derived a general and practicalmodel for dissipation and vertically averaged wave orbital velocities applica-ble to very dense canopies. Extensions were developed for two-layer canopies(Weitzman et al., 2015) and random waves, and tested or calibrated againstobserved wave attenuation (Jadhav et al., 2013; Lowe et al., 2007). Zelleret al. (2015) found that major model simplifications can often be justified:in many natural canopies, acceleration-dependence of drag (resulting fromFroude-Krylov forces and added mass, e.g. Sumer and Fredsøe, 1997) andvertical turbulent mixing can both be neglected at wave frequencies. Build-ing on the approach of Zeller et al. (2015), we examine the properties of asimple model for wave orbital velocities, and test the model against observa-tions within natural near-bed canopies of mangrove pneumatophores.

Previous laboratory experiments have quantified enhanced dissipation(Pujol et al., 2013a) and reduction of wave orbital velocities (Lowe et al.,2005; Luhar et al., 2010; Pujol et al., 2013b; Weitzman et al., 2015) by dragwithin artificial canopies. In natural canopies, many stems, each having itsown unique size and shape, often create substantial vertical variability. Suchvertical variability is here quantified using recently developed photogrammet-ric techniques, which measure the three-dimensional geometry of all stemswithin sampled regions of canopy (Lienard et al., 2016). Combining datafrom photogrammetric surveys with wave orbital velocities measured withinnatural pneumatophore canopies, we test a simple vertically resolved modelderived from the equations of Zeller et al. (2015). Modeling is then appliedto analyze the ability of Mekong Delta pneumatophore canopies to shelteronshore regions from incoming waves.

The model is outlined in Section 2, in turn considering within-canopyvelocity (Section 2.1) and dissipation (Section 2.2). The field site, instru-mentation, and data analysis are outlined in Section 3. Results presented inSection 4, and summarized in Section 5.

2. Theory

2.1. Vertical Variability of Velocity Induced by Canopy Drag

The vegetation canopy is described by statistics n =number of stems persquare meter, d =mean stem diameter, a =cross-sectional area of vegetation

4

Page 5: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

blocking the flow per cubic meter of canopy (units m�1), and � =proportionof volume occupied by solid stems. For roughly cylindrical vertical stems,a ⇡ nd and � ⇡ (⇡/4)nd2. Here n is a function of elevation z above the bed,because the number of stems usually decreases from a maximum at the bedto zero immediately above the elevation hv of the highest stem (a, d and �also vary with elevation).

As noted by Zeller et al. (2015), for many natural canopies � ⌧ 1, and atwave frequencies vertical mixing of momentum and acceleration-dependenceof canopy drag forces are negligible (assumptions required for these simplifi-cations are noted below and in Appendix A). For weakly nonlinear waves(|u| ⌧ c, where c =wave phase speed), the horizontal momentum equa-tion (e.g. eq.1.25 of Zeller et al., 2015, with wave-induced momentum fluxneglected) then reduces to

@u

@t+ FD +

1

@p

@x= 0, (1)

where t =time, u =water velocity in horizontal direction x, ⇢ =water den-sity, p =pressure, the canopy drag force per kilogram of water is

FD =CD

2a|u|u, (2)

CD is the drag coe�cient for flow past a stem (here CD ⇡ 2, Section 4),and |u| is the magnitude of u. Therefore, acceleration [first term of (1)]results from vegetation drag (second term), and lateral pressure gradients(third term). Equation (1) is obtained by averaging over a horizontal regionencompassing many stems (Zeller et al., 2015), but extending substantiallyless than one wavelength in any direction. For a velocity ub above the canopy,(1) reduces to

@ub

@t+

1

@⇢

@x= 0. (3)

Let

Tf =2

CDa|u|. (4)

Substituting FD = u/Tf [which follows from (2) and (4)] into (1) shows thatTf is a frictional timescale; if pressure forcing were negligible then drag wouldcause an initial velocity to decay over time of order Tf . Depth-uniform hori-zontal pressure gradients are eliminated by subtracting (3) from (1), leading

5

Page 6: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

to@(u� ub)

@t+

u

Tf= 0. (5)

From (5),Typical magnitude of u� ub

Typical magnitude of u= O(✏f ), (6)

where the ordering parameter

✏f =Tw

2⇡Tf=

CDa|u|Tw

4⇡, (7)

Tw =typical wave period, and O(✏f ) denotes a term of order ✏f . Therefore,if the frictional timescale Tf is much longer than the wave period Tw, drag isinsu�cient to generate a large di↵erence between within- and above-canopyvelocities.

To clarify the assumptions underlying (1), let

✏⌫ =Cf |u|Tw

2⇡hv. (8)

Perturbation analysis (Appendix A) resembling the scaling of Zeller et al.(2015) establishes (1) from more general equations [e.g. from equation (4) ofLKM05] given u ⌧ c, � ⌧ 1 and ✏⌫ ⌧ max{✏�1

f , 1}. Under these conditionsvertical mixing [neglected in (1) and (5)] is negligible through most of thecanopy, although mixing remains significant in a thin neighborhood of thebed called the Wave Bottom Boundary Layer (WBBL, Mei, 1989).

For simplicity, we solve (5) for a time series discretely sampled at timestj = j�t, with j between �N and N . Velocities are represented by theFourier series

u =NX

j=�N

huiTj ei2⇡t/Tj , (9)

ub =NX

j=�N

hubiTj ei2⇡t/Tj , (10)

where, for any variable �, h�iTj denotes the complex amplitude at period Tj =(2N + 1)�t/j. To derive a spectral model, we set |u| to the constant value(8/⇡)1/2urms in (4), where urms is the root-mean-squared value of u. Here

6

Page 7: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

the factor (8/⇡)1/2 was chosen to give the correct mean value for dissipation,given a Gaussian velocity distribution. Similar approximations have beenused for WBBL models (Tolman, 1994), for spectral models of forces oncylinders (Borgman, 1967), for simulating forces on saltmarsh vegetation(Mullarney and Henderson, 2010) and for simulating wave dissipation in coraland vegetation canopies (Jadhav et al., 2013; Lowe et al., 2007). From (5),(9), and (10), within- and above-canopy velocities are related by

huiTj = �jhubiTj , (11)

where the transfer function

�j =1

1� i✏j, (12)

and the value of ✏f for the jth Fourier component is [c.f. (7)]

✏j =Tj

2⇡Tf=

CDa|u|Tj

4⇡. (13)

For small ✏j, (12) predicts an O(✏j) phase lead (within-canopy velocity aheadof overlying velocity), and a smaller O(✏2) reduction in velocity amplitude.From (13), ✏j is proportional to wave period Tj, so both phase lead andamplitude attenuation increase with increasing period.

The solution (12) attributes all amplitude attenuation to canopy drag.However, even in the absence of drag, wave velocities decline with increasingdepth unless the wavelength greatly exceeds the depth (i.e. unless waves arein shallow water, Mei, 1989). Frictionless linear wave theory predicts a ratio⇣ = cosh(kzl)/ cosh(kzu) between velocities at lower and upper elevations zland zu, where k = 2⇡/wavelength is calculated from the dispersion relation(2⇡/Tj)2 = gk tanh(kh), and g =9.80ms�2 (Lowe et al., 2007). For thecanopies considered here, such frictionless within-canopy depth attenuationwill prove insignificant (⇣ ⇡ 1) for wave frequencies below about 0.3Hz.At higher frequencies, ⇣ will drop substantially below 1, but for these cases�j ⇡ 1. Therefore, a transfer function accounting for both frictional andfrictionless vertical attenuation is

�j =⇣

1� i✏j. (14)

This prediction will be compared with observations.

7

Page 8: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

2.2. Simulated DissipationLetting x be the wave propagation direction and neglecting bottom fric-

tion and directional spread, the depth-integrated wave energy balance atperiod Tj is

@ETjcg,j

@x= �

Z h

0

ETj dz (15)

where ETj is the spectral density of depth-integrated wave energy at periodTj, cg,j is the associated group velocity, and ETj is the spectral density ofdissipation by vegetation drag (a function of elevation). Let �Tj(X, Y ) bethe cross spectrum at period Tj between any two time series X and Y (so�Tj(X,X) is a power spectrum). The dissipation of wave energy by vegeta-tion drag is the mean of uFD, so from (2) and (11) the spectral density ofdissipation is [c.f. equation 27 of Lowe et al. (2007) and equation 7 of Jadhavet al. (2013)]

ETj =Cd

2a|u||�|2�Tj(ub, ub), (16)

where the velocity ub immediately above the pneumatophores will be approx-imated by the velocity predicted at the bed by frictionless linear wave theory(as is appropriate for near-bed canopies). Expressions (13) – (14) for � de-pend on within-canopy |u|, which itself depends on �. To obtain an explicitpredictive model for wave attenuation, an approximation for � in terms ofabove-canopy velocity |ub| is now derived. Since �j is a smooth function ofTj, assuming a narrow-banded spectrum with peak period Tw the transferfunction is approximated by a single value �0 applicable to the full spectrum(this approach resembles Lowe et al., 2007). Substituting |�0| = |u|/|ub| into(13) and (14) then yields

�0 =1

1� i✏0|�0|, (17)

where the dimensionless damping parameter

✏0 =CDa|ub|Tw

4⇡(18)

di↵ers from ✏f in that the above-canopy velocity ub appears in (18) whereasthe within-canopy velocity u appears in (7). For ✏0 = 0 this yields �0 = 1,and for ✏0 > 0 some algebra yields

|�0|2 =(1 + 4✏20)

1/2 � 1

2✏20. (19)

8

Page 9: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

The depth-average of |�0| was denoted ↵w by LKM05. The model consideredhere does not correspond to any of the simple limits (canopy-independent,inertia-dominated, or unidirectional) considered by LKM05. Consistent withthe general analysis of LKM05, the orbital displacement (denoted A1

rms byLKM05 and proportional to |u|Tw here) is the sole hydrodynamic variablecontrolling |�0|. However, |�0| here depends on only ✏0, which di↵ers bya factor of order �1/2CD/(2⇡3/2) from the parameter A1

rms/S considered byLKM05, where the spacing between stems S ⇡ n�1/2.

From (16), (18) and (19)

ETj = �2⇡�Tj(ub, ub)

Tw, (20)

where the dimensionless dissipation

� =[(1 + 4✏20)

1/2 � 1]3/2

23/2✏20. (21)

Now 4⇡� is the ratio between the dissipation in one wave period and theabove-canopy energy kinetic energy (both expressed per kilogram of water).For su�ciently sparse canopies (✏0 ⌧ 1), � ⇡ ✏0, and since ✏0 is proportionalto a (18), increasing vegetation density is associated with increasing dissipa-tion (Figure 1). Expression (20) with � = ✏0 could alternatively have beenderived by simply neglecting all within-canopy attenuation of velocity (aswas done, for example, by Dalrymple et al., 1984), so this limit is called the‘unattenuated dissipation model’. As ✏0 increases above about 0.4 (verticaldotted line, Figure 1), dissipation departs from the unattenuated prediction,and � reaches a maximum of 1/2 when ✏0 = 21/2. Further increase in ✏0reduces � [� ⇡ ✏

�1/20 for ✏0 � 1]. Such complex behaviour is possible owing

to reduced wave orbital velocities, and resulting reduced dissipation, withinvery dense canopies. Lowe et al. (2007) previously derived a more generalmodel for this reduction in wave dissipation resulting from reduced within-canopy wave orbital velocities. The analysis here is novel in deriving ananalytic expression, which highlights the importance of ✏0 for the limit oflow � and negligible vertical mixing.

Solution of the model (15), (18), (20), and (21) for arbitrary incidentwaves propagating into a forest with variable geometry and bathymetry re-quires numerical integration. However, for ✏0 ⌧ 1 and constant depth theunattenuated model yields a simple analytic solution. From (18), (20), and

9

Page 10: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

� ⇡ ✏0, the depth-integrated dissipation

Z h

0

E! dz =CD�|ub|

2�Tj(ub, ub), (22)

where

� =

Z h

0

a dz. (23)

Substituting (22) and the linear wave theory resultETj = (h/2)[1 + sinh(2kh)/(2kh)]�Tj(ub, ub) into (15) and solving yields

ETj =E0

(1 + x/x0)2, (24)

where the dissipation length scale

x0 =2hcg,j[1 + sinh(2kh)/(2kh)]

CD�|u0|(25)

is the distance waves propagate into the marsh before their amplitude ishalved by dissipation, E0 is the energy at x = 0, and |u0| is |ub| at x = 0.This closely resembles a solution found by Dalrymple et al. (1984), heresimplified for a near-bed canopy.

3. Field Site, Instrumentation, and Data Analysis

We present measurements of water velocity and vegetation geometry ob-tained along the forested, ocean-exposed coast of Cu Lao Dung Island in theSouthern Mekong Delta. Within 0.8m of the bed, dense canopies of pneu-matophore stems grow upward from the buried lateral roots of the dominantSonneratia caseolaris mangroves. Two control deployments C1 and C2 wereconducted on the unvegetated tidal flats, o↵shore of the forest edge. The re-maining deployments P1 –P5 sampled several locations within the forest, andseveral elevations within the pneumatophore canopy. Deployments C1 –C2and P2 –P4 were conducted within 15m of forest-edge location N9�29.5090’,E 106�14.6377’, near the Southwest corner of Cu Lao Dung Island. Deploy-ments P1 and P5 were conducted at the forest-edge location N9�33.9296’,E 106�17.5562’, near the Northeast corner of the Island. The pneumatophorecanopies near the forest edge, sampled during P1 –P5, were usually more

10

Page 11: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

10-1 1 10

1 10

2

ϵ0

10-1.5

10-1

10-0.5

1

χ

Figure 1: Dimensionless dissipation � [thin black curve, see equations (20) – (21)] versusdimensionless damping parameter ✏0, showing maximum dissipation for order-one ✏0. Solidgrey curve: approximation � = ✏0, valid for ✏0 ⌧ 1. Dashed grey curve: approximation

� = ✏�1/20 , valid for ✏0 � 1. Dotted black vertical line: ✏0 =0.4.

Figure 2: [This is the only figure to be color in print version]Image (a) and three-dimensional reconstruction (b) of low-density pneumatophore canopy P2. The greyquadrat, marking one square meter, was removed before the incomming tide flooded thesampling region. The Vector (yellow square) and Vectrino (orange triangle) current metersmeasured above- and within-canopy velocity. The aquadopp current meter (red diamond)was not used here.

11

Page 12: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

dense (larger n and d) than canopies farther into the forest interior. Addi-tional deployment details are given by Mullarney et al. (2016); Norris et al.(2016) [this issue].

During each deployment, a pair of vertically displaced current meters si-multaneously measured velocities at ‘lower’ (0.019m zl 0.38m) and ‘up-per’ (0.3m zu 0.93m) elevations (Columns 1 – 4, Table 3). During P1 –P5, the lower and upper instruments respectively measured velocity withinand above the pneumatophore canopy. Deployment names indicate the rela-tive canopy density at the lower instrument elevation, with the lowest densityfor P1 and the highest density for P5. Instruments deployed in the low den-sity case P2 are shown in Figure 2a. ‘Deployments’ P1 and P5 actually referto a single case, where a vertical stack of three instruments (one above thecanopy and two within the canopy) were deployed for 20 hours, and twodi↵erent time series were used to obtain maximum sampling duration (one4.8-hour time series was was centered on the first high tide observed duringthis deployment, and the second 4.5-hour time series was centered on thesecond high tide; spectra from these two time series we averaged to yield theresults presented, although similar results are obtained if these time series areanalyzed separately). The upper and middle instruments were compared in‘Deployment P1’, whereas the upper and lower instruments were comparedin ‘Deployment P5’. For remaining cases (C1 –C2, P2 –P4), measurementcommenced soon after the upper instrument was submerged by the risingtide, and instruments were removed within a few hours, limiting samplingdurations to 157 minutes. Except for a few brief stoppages required fortechnical reasons, these time series were continuous. Deployments P3 and P4were conducted at the same horizontal location on successive days, whereaseach of C1, C2, and P2 was conducted at a di↵erent location. The lowerinstrument was always a Nortek Vectrino profiler, whereas the upper instru-ment was either a Vectrino profiler (deployments C1, P1, P5) or a NortekVector (C2, P2, P3, P4). Vectrino profilers measured velocity at 50Hz everymillimeter along a 35mm profile. For deployments C1, P1, and P3, data fromeither end of each Vectrino profile were discarded, and velocities from the cen-tral 15mm were averaged to yield a single 50Hz time series for each Vectrino.For deployments C2, P2, P4, and P5, where the mid profile for the near-bedVectrino was sometimes within 2 cm of the bed, velocities from to highest5mm of the profile were averaged, to minimize the influence of the WBBL(in these cases Vectrinos provided acoustic tracking of the bed, from whichthe mean sampling elevation zl was determined). Vectors measured point

12

Page 13: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

velocity at 32Hz, which was linearly interpolated to 50Hz for comparisonwith Vectrino data. All Vectrinos were cabled to a single computer, whereasVectors logged data internally. To synchronize Vectors with Vectrinos, Vec-tor clock times were adjusted (usually by < 1 s) to ensure minimal phaseshift between vertically displaced instruments at high (> 0.5Hz) frequencies(these clock adjustments had minimal e↵ect on phases at longer periods).Small (< 5 degree) errors in rotation about the vertical axis were removedby optimizing the correlation between upper and lower velocity time series.Results are presented for horizontal velocity u in the mean wave direction(roughly onshore, and defined as the principal axis of measured 0.1 – 0.8Hzvelocity, Kuik et al., 1988). Low (< 70%) correlation data and spikes werereplaced by linear interpolation.

Photographic surveys of pneumatophore canopies were conducted at lowtide, within one day of corresponding current meter deployments. Pho-tographs obtained from many (55 – 387) angles around 1m2 quadrats cen-tered on the current meters were used to reconstruct a three-dimensionalpoint cloud using the open-source photogrammetric software VisualSFM(Figure 2b). The diameter and location of every pneumatophore within eachquadrat was estimated every 5mm along vertical profiles extending fromabout z =0.03m to the top of the canopy, using the point cloud and thetechniques developed by Lienard et al. (2016). For each of deployments P1 –P5, vertical profiles of vegetation statistics n, d, a and � were calculated fromthese data.

To test (14), empirical transfer functions between the upper velocity uu

and the lower velocity ul were estimated. The variance of upper and lowervelocities is quantified as a function of wave period by the power spectra�Tj(uu, uu) and �Tj(ul, ul). The cross-spectral matrix at period Tj betweenthe upper and lower velocities is

M(Tj) =

�Tj(uu, uu), �Tj(uu, ul)�Tj(ul, uu), �Tj(ul, ul)

�. (26)

The (complex) dominant eigenvector of M(Tj), denoted (uu, ul), is the dom-inant EOF for period-Tj fluctuations in the two velocities (Henderson et al.,2001). An empirical transfer function was estimated from this dominantEOF as

�Tj =ul

uu. (27)

13

Page 14: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

The magnitude and argument of this complex transfer function are respec-tively called the ‘Gain’ and the ‘Phase’ [the phase simply equals the phaseof the cross spectrum �Tj(uu, ul)]. Since the estimated transfer function ismeaningful only when upper and lower layers are coherent, we also presentthe squared coherence |�Tj(uu, ul)|2/[�Tj(uu, uu)�Tj(ul, ul)]. For each de-ployment, spectra were estimated using 50%-overlapping, hanning-windowedtime series segments (Welch, 1967), yielding > 50 degrees of freedom.

To compare theoretical and empirical transfer functions [(14) and (27)],we write ✏j = ↵|u|Tj/(4⇡), and for each deployment choose the single ↵

value that minimizes the squared error |�Tj � �Tj |2 summed over all coherentfrequencies (coherence2 > 0.5). Here |u| is calculated as (8/⇡)1/2urms, whereurms =root-mean-squared lower velocity. From (4) and (13) ↵ = CDa, so aplot of ↵ values obtained from multiple deployments against correspondinga values should show a positive correlation, with zero o↵set and slope equalto CD. Laboratory experiments with oscillating cylinders indicate that CD

depends on the stem Reynolds number Res = 2urmsd/⌫m and the Keulegan-Carpenter numberKC = 2urmsTw/d, where ⌫m =kinematic viscosity of waterand Tw is a peak period. For Re = O(103) and KC =50 – 300, experimentssuggest CD =1– 3 (e.g. Sumer and Fredsøe, 1997).

To examine whether drag-induced reduction of within-canopy wave or-bital velocities substantially reduced dissipation, ✏0 was calculated from (18)and observations. The percentage overestimation in dissipation that wouldresult from using the unattenuated model is (✏0/�� 1)⇥ 100%.

Label Date Duration zl, zu a Tp uu ↵ ✏0(minutes) (m) (m�1) (s) (ms�1)

C1 9/27/2014 43 0.10, 0.30 0 2.0 0.29 0.012 0C2 3/11/2015 45 0.019, 0.45 0 6.7 0.53 0.10 0P1 3/13/2015 553 0.38, 0.81 0.14 2.2 0.28 0.16 0.015P2 3/11/2015 37 0.028, 0.60 0.57 6.7 0.22 0.42 0.13P3 3/8/2015 157 0.19, 0.93 0.58 5.9 0.20 1.0 0.17P4 3/7/2015 129 0.021, 0.4 1.45 2.7 0.17 2.8 0.18P5 3/13/2015 553 0.021, 0.81 1.78 2.2 0.28 4.2 0.38

Table 1: Summary of deployments. zl, zu =elevations of lower and upper velocity mea-surements; vegetation area a evaluated at elevation zl; Tp =peak period of upper velocityspectrum; uu =root-mean-square upper velocity magnitude; ↵ =fitting parameter (theo-retical value CDa); ✏0 =dimensionless damping parameter.

14

Page 15: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

0 30 60 90

n (m-2

)

0

0.1

0.2

0.3

0.4

0.5

0.6

z (m

)

a

0 0.01 0.02 0.03

d (m)

b

0 1 2

a (m-1

)

c

0 0.02 0.04

φ

d

Figure 3: Vertical profils of vegetation statsitics. Panel a: number of stems per squaremeter n. Panel b: mean stem diameter d. Panel c: frontal area a. Panel d: solid volumefraction �. Thin black curves: canopy measured at location of deployments P1 and P5.Thick light grey: location of P2. Dashed dark grey: location of P3 and P4.

4. Results

During the seven deployments, a ranged from 0 to 1.78m�1, peak waveperiods ranged from 2.0 to 6.7 s, and rms velocity magnitudes ranged from0.17 to 0.53ms�1 (Table 3). During P1 –P5, stem Reynolds numbers rangedfrom 3.3 ⇥ 103 to 5.4 ⇥ 103, and Keulegan-Carpenter numbers ranged from91 to 290.

Pneumatophore canopies were about 0.6m high, with 45 – 84 stems m�2,frontal area a 1.8m�1, mean stem diameters d 0.028m, and solid volumefraction � 0.030 (Figure 3). The lowest density canopy was measured atthe location of deployment P2 (light grey curves, Figure 3, and pictured inFigure 2). Although profiles of stem number and diameter di↵ered betweenthe location of deployments P1 and P5 and the location of deploymentsP3 and P4, the two locations showed similar profiles of a (solid black anddashed dark grey curves, Figure 3). In all cases, � was su�ciently low thatwake interference between stems (Tanino and Nepf, 2008) and acceleration-dependent forces were negligible.

During deployment C1, short period, moderate energy waves were ob-served (Table 3). As in all deployments, the velocity power spectrum wasdominated by 0.1 – 1Hz (hereafter ‘incident frequency’) waves (Figure 4a).Upper and lower velocities were nearly equal at most incident frequencies,with similar power spectra (Figure 4a), strong coherence (Figure 4b), gain

15

Page 16: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

10-4

10-3

10-2

Po

wer

(m

2s-1

)a

0

0.5

1

Co

her

ence

2 b

0

0.5

1

Gai

n

c

0.01 0.03 0.10 0.30 1.00

Frequency (Hz)

-20

0

20

Ph

ase

(deg

.) d

Figure 4: Comparison of upper (elevation z =0.30m) and lower (z =0.10m) velocitiesmeasured over unvegetated tidal flats, Deployment C1, 27 September 2014. (a) powerspectra of upper (grey) and lower (black) velocities. (b – d): Squared coherence (b), gain(c, < 1 indicates lower velocity smaller magnitude), and phase (c, positive indicates lowervelocity leading) between upper and lower velocities. In (c) – (d), black (grey) dots indicatefrequencies with squared coherence greater (less) than 0.5, and grey curves indicate fittedtheoretical transfer function (14). In (c), black dashed curve, in this case indistinguishableform grey curve, indicates theoretical depth attenuation predicted by non-dissipative linearwave theory.

16

Page 17: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

10-3

10-2

10-1

Po

wer

(m

2s-1

)

0

0.5

1C

oh

eren

ce2

0

0.5

1

Gai

n

0.01 0.03 0.10 0.30 1.00

Frequency (Hz)

-20

0

20

Ph

ase

(deg

.)

Figure 5: As figure 4, but for deployment C2, over unvegetated tidal flats, 11 March 2015,upper and lower elevations z =0.45 and 0.019m.

usually near 1 (Figure 4c) and phase near zero (Figure 4d). However, at thehighest frequencies (> 0.5Hz) the gain dropped, consistent with the depthattenuation predicted by frictionless linear wave theory (black dots and blackdashed curve, Figure 4c). At low frequencies (< 0.1Hz), energy levels werelow (Figure 4a) and upper and lower velocities were only weakly coherent(Figure 4b). Since surface wave velocities are essentially depth-uniform atthese frequencies, these vertically incoherent low frequency motions werelikely not surface waves, but may have been turbulent eddies. Fitting themodel (14) to the observed transfer function (27) yields a fitted transfer func-tion near 1 (grey curves, Figure 4c,d) and a fitting parameter ↵ near zero(eighth column, table 3), as expected in the absence of vegetation.

More energetic, longer period waves were observed during C2 (Table 3).As in deployment C1, upper and lower velocities were nearly equal at 0.1 –

17

Page 18: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

0.5Hz, with some depth attenuation resembling the linear theory predictionat higher frequencies (Figure 5a – d). At lower frequencies (< 0.1Hz) a smallphase shift was observed, with the lower velocity leading by about 5�. Gainalso dropped below 1 at about 0.03Hz. A gain< 1 and a phase lead up to25� are expected within the WBBL, and the lower current meter was nearthe bed (zl =0.019). Although the incident-frequency WBBL was likely toothin to be measured by the lower current meter, boundary layer thicknessincreases with wave period (Mei, 1989), so the gain and phase observed atfrequencies < 0.01Hz may indicate the outer edge of the infragravity wavebottom boundary layer. Fitting the theoretical transfer function yields � near1, and a fitting parameter ↵ that is small, although not as small as in caseC1 (Table 3). Regardless of the small phase shift at low frequencies, bothC1 and C2 indicate that, in the absence of vegetation, depth-dependenceat energetic incident frequencies was minimal, except for frictionless depthattenuation at the highest frequencies.

In contrast to deployments C1 –C2 on the flats, clearer depth-dependencewas observed within pneumatophores. Deployment P2, in a relatively lowdensity canopy (Table 3), showed a phase shift of about 10� and 0.1Hz (Fig-ure 6d). Consistent with theory, phase decreased with increasing frequency.The model (14) provides a reasonable fit to observations (compare black dotsand grey curves, Figure 6c,d), with a significantly non-zero fitting parame-ter ↵ (Table 3). Nevertheless, departures of the transfer function from thefrictionless value ⇣ were small and the damping parameter ✏0 = 0.13 wassubstantially less than 1. Results were similar for deployment P3 (Figure 7),with clearer frictionless depth attenuation at high frequencies resulting froma relatively large separation of zu and zl. The greatest attenuation and phaseshift were observed in the high density canopy of P5 (Figure 8, note the ex-tended vertical axis range of Figure 8d). At low frequencies (< 0.03Hz), gainapproached zero and phase approached 90� (Figure 8c,d), consistent with the✏j ! 1 limit of (14), although associated coherence was low (Figure 8b).The estimated ✏0 =0.38 was the largest for any deployment. Therefore, ob-servations spanned the region to the left of the vertical black dotted linein Figure 1, where the unattenuated model provides a good approximation(solid grey and black curves match). For ✏0 =0.38, the unattenuated modelis expected to overestimate dissipation by 20%.

As expected given the theoretical relationship ↵ = CDa, values of ↵obtained by fitting transfer functions for the seven deployments were cor-related with corresponding a values (Figure 9). The best-fit slope suggests

18

Page 19: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

10-3

10-2

10-1

Po

wer

(m

2s-1

)

a

0

0.5

1

Co

her

ence

2 b

0

0.5

1

Gai

n

c

0.01 0.03 0.10 0.30 1.00

Frequency (Hz)

-20

0

20

Ph

ase

(deg

.) d

Figure 6: As figure 4, but for deployment P2, vegetation density a =0.57, 11 March 2015,upper and lower elevations z =0.60 and 0.028m.

19

Page 20: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

CD =2.1, comparable to values for oscillating cylinders at these Reynoldsand Keulegan-Carpenter numbers. As expected given dominance by canopydrag (rather than the vertical mixing responsible for WBBL vertical struc-ture), there is no tendency for velocities measured within 3 cm of the bed(black symbols, Figure 9) to yield di↵erent ↵ values than velocities measuredat higher elevations (grey symbols).

For the observed small values of ✏0, the unattenuated model (25) pro-vides an estimate of the distance x0 waves can propagate over uniform pneu-matophore canopies before their amplitude is halved. Trapezoidal integra-tion of measured a profiles yields � (see 23) of 0.095 (location of P2), 0.29(location of P3, P4) and 0.28 (location of P1, P5). Here, we focus on dis-sipation by pneumatophores, rather than tree trunks, because a values aremuch higher within pneumatophore canopies. Given CD =2, depth h =2m,and observed wave conditions, associated wave decay distances were 510m(P2), 180m (P3), 200m (P4), and 180m (P1, P5). Although these decaydistances are large, predictions are sensitive to the assumed water depth; fora depth of 1m, decay distances are reduced to 180m (P2), 65m (P3), 74m(P4), and 48m (P1, P5).

5. Summary

On Cu Lao Dung Island in the Mekong Delta, wave orbital velocities mea-sured within pneumatophore canopies were roughly consistent with a simplemodel (Zeller et al., 2015) that neglects vertical mixing and acceleration-dependent drag. Frequency dependence matched an analytic solution pre-dicted using a linearized drag law. When solving for the dissipation of allwave energy (rather than frequency dependence) the model considered hereis nonlinear, but su�ciently simple that an analytic solution relates within-canopy flow to above-canopy flow.

In low density canopies, modeled dissipation is proportional to canopydensity, measured by a. Conversely, in su�ciently high density canopies,increasing density can reduce near-bed flows enough to reduce dissipation(dissipation proportional to a�1/2). The transition between the low- and high-density regimes is controlled by a dimensionless parameter ✏0 (18), which is afunction of both canopy geometry and hydrodynamic conditions; dissipationis maximum at the transition between low- and high-density regimes, whichoccurs near ✏0 =1.4 (Figure 1). The pneumatophore canopies and hydro-dynamic conditions we measured on Cu Lao Dung Island were in the low

20

Page 21: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

10-3

10-2

10-1

Po

wer

(m

2s-1

)

a

0

0.5

1

Co

her

ence

2 b

0

0.5

1

Gai

n

c

0.01 0.03 0.10 0.30 1.00

Frequency (Hz)

-20

0

20

Ph

ase

(deg

.) d

Figure 7: As figure 4, but for deployment P3, vegetation density a =0.58, 8 March 2015,upper and lower elevations z =0.93 and 0.19m.

21

Page 22: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

10-4

10-3

10-2

Po

wer

(m

2s-1

)

a

0

0.5

1

Co

her

ence

2 b

0

0.5

1

Gai

n

c

0.01 0.03 0.10 0.30 1.00

Frequency (Hz)

-90

0

90

Ph

ase

(deg

.) d

Figure 8: As figure 4, but with extended y axis range, and for deployment P5, vegetationdensity a =1.78, 13 March 2015, upper and lower elevations z =0.81 and 0.021m.

22

Page 23: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

0 0.5 1 1.5 2

a (m-1

)

1

2

3

4

α (

m-1

)

Figure 9: Fitting parameter ↵ for hydrodynamic model versus measured vegetation den-sity parameter a. Dashed line indicates theoretical relationship ↵ = CDa with best-fitCD =2.1. Deployment C1 grey circle, C2 black circle [partially obscured, near (0,0)], P1grey diamond, P2 black square, P3 grey triangle, P4 black triangle, P5 black diamond.Black symbols: lower velocity measured at elevation zl < 0.03m above bed (grey symbols:zl > 0.03m).

23

Page 24: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

density regime (✏0 0.38). This finding is consistent with the increase inwithin-canopy turbulent dissipation with increasing a found by Norris et al.(2016) [this issue]. Nevertheless, ✏0 was not much less than 1, so occasionaldeparture from the low density regime remains likely. For one observed case(P5), a doubling of wave period and wave height, plausible in storm condi-tions, would yield ✏0 = 1.5, at the boundary between low- and high-densityregimes. In this case, using frictionless linear theory to estimate within-canopy velocity would yield wave dissipation estimates triple the true value.Elevated dissipation within pneumatophores may enhance sediment mobi-lization (Norris et al., 2016, this issue), whereas sheltering of the bed at highcanopy densities could create conditions for rapid deposition. Rapid depo-sition can bury pneumatophores, reducing mangrove vigor or even causingdeath (Ellison, 1999; Mo↵ett et al., 2015; Nardin et al., 2016). Therefore,the tendency for observed pneumatophore canopies to approach, but seldomexceed, the density of maximum dissipation may be adaptive.

By enhancing dissipation at the forest edge, pneumatophores reduce theheight of waves propagating into the forest interior. Theory for low canopydensity (✏0 ⌧ 1), resembling the model of Dalrymple et al. (1984), predictsstrong depth-dependence to this sheltering e↵ect, qualitatively consistentwith previous observations (Mazda et al., 2006). For observed waves andpneumatophore canopies, simulations predict that waves dissipated substan-tially as they propagated tens of meters in 1m water depth, or hundreds ofmeters in 2m depth (propagation distances would be shorter under higherwave energies). However, the observations reported here were collected nearthe forest fringe, where canopies were relatively dense (Norris et al., 2016, thisissue). Therefore, lower canopy densities might lead to reduced dissipationin the forest interior.

Acknowledgements

During fieldwork in Vietnam, vital logistical support was provided byHuong Phuoc Vu Luong, Charles Nittrouer, Richard Nguyen, Andrea Ogston,and Daniel Culling. Further assistance at the field site was provided byDean Sandwell, Aaron Fricke, Daniel Culling, Sergio Fagherazzi, WilliamNardin, Xuan Tien Nguyen Vinh, and Hoang Phong Nguyen. Fieldwork wasfunded by the US O�ce of Naval Research, Littoral Geosciences grant num-ber N000141410112, and the O�ce of Naval Research Global grant numberN62909-14-1-N028.

24

Page 25: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

Appendix A. Perturbation expansion

In the limit of small � (denoted �p by LKM05), subtracting our equation(3) from equation (4) of LKM05 yields

@ud

@t+

u

Tf� ⌫

@2ud

@z2= 0, (A.1)

where the defect velocity ud = u � ub and the eddy viscosity ⌫ = Cf |ud|hv.Dimensionless variables, denoted by ⇤, are defined by

t =Twt⇤2⇡

, (A.2)

z =hvz⇤. (A.3)

Consider separately the cases ✏f 1 and ✏f > 1.

Appendix A.1. Case ✏f 1

For moderate or weak attenuation of within-canopy velocity, define di-mensionless variables by

u =u0u⇤, (A.4)

ud =✏fu0ud⇤, (A.5)

⌫ =Cf✏fu0hv⌫⇤. (A.6)

Now (A.1) becomes

@ud⇤

@t⇤+ u⇤ � ✏⌫✏f⌫⇤

@2ud⇤

@z2⇤= 0, (A.7)

so mixing is negligible if ✏f✏⌫ ⌧ 1.

Appendix A.2. Case ✏f > 1

For the case of large attenuation, we rescale

u1 =u0u1⇤, (A.8)

u =u0u⇤

✏f, (A.9)

ud ⇡� u1 = �u0u1⇤, (A.10)

⌫ =Cfu0hv⌫⇤. (A.11)

25

Page 26: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

Now (A.1) becomes

�@u1⇤

@t⇤+ u⇤ � ✏⌫⌫⇤

@2ud⇤

@z2⇤= 0, (A.12)

so mixing is negligible if ✏⌫ ⌧ 1.

References

Anthony, E. J., Brunier, G., Besset, M., Goichot, M., Dussouillez, P., andNguyen, V. L. (2015). Linking rapid erosion of the Mekong River Delta tohuman activities. Scientific Reports, 5:doi:10.1038/srep14745.

Bao, T. (2011). E↵ect of mangrove forest structures on wave attenuation incoastal Vietnam. Oceanologia, 53(3):807–818.

Borgman, L. (1967). Spectral analysis of wave forces on piling. J. Waterwaysand Harbors Div., ASCE, 93:129 – 156.

Bouma, T., van Duren, L., Temmerman, S., Claverie, T., Blanco-Garcia,A., Ysebaert, T., and Herman, P. (2007). Spatial flow and sedimentationpatterns within patches of epibenthic structures: combining field, flumeand modelling experiments. Cont. Shelf Res., 27:1020 1045.

Dalrymple, R. A., Kirby, J., and Hwang, P. (1984). Wave di↵raction dueto areas of energy dissipation. Journal of Waterway, Port, Coastal, andOcean Engineering, 110:67 – 79.

Ellison, J. C. (1999). Impacts of sediment burial on mangroves. MarinePollution Bulletin, 37(812):420 – 426.

Furukawa, K., Wolanski, E., and Mueller, H. (1997). Currents and sedimenttransport in mangrove forests. Estuar. Coast. Shelf Sci., 44:301 310.

Giri, C., Ochieng, E., Tieszen, L., Zhu, Z., Singh, A., Loveland, T., Masek,J., and Duke, N. (2011). Status and distribution of mangrove forests ofthe world using earth observation satellite data. Glob. Ecol. Biogeogr.,20:154 – 159.

Greenberg, R., Maldonado, J., Droege, S., and Mcdonald, M. (2006). Tidalmarshes: a global perspective on the evolution and conservation of theirterrestrial vertebrates. Bioscience, 56:675 – 685.

26

Page 27: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

Henderson, S. M., Elgar, S., and Bowen, A. (2001). Observations of surfbeat propagation and energetics. In Proceedings of the 27th InternationalConference on Coastal Engineering, pages 1412–1421. ASCE.

Horstman, E. M., Dohmen-Janssen, C. M., Narra, P. M. F., van den Berg,N. J. F., Siemerink, M., and Hulscher, S. J. M. H. (2014). Wave attenua-tion in mangroves: A quantitative approach to field observations. CoastalEngineering, 94:47 – 62.

Jadhav, R. S., Chen, Q., and Smith, J. M. (2013). Spectral distribution ofwave energy dissipation by salt marsh vegetation. Coastal Engineering,77:99–107.

Kuik, A., Van Vledder, G. P., and Holthuijsen, L. (1988). A method forroutine analysis of pitch-and-roll buoy wave data. J. Phys. Oceanogr.,18:1020–1034.

Lienard, J., Lynn, K., Strigul, N., Norris, B., Gatziolis, D., Mullarney, J. C.,Bryan, Karin, R., and Henderson, S. M. (2016). E�cient three-dimensionalreconstruction of aquatic vegetation geometry: Estimating morphologicalparameters influencing hydrodynamic drag. Estuarine, Coastal, and ShelfSciences, 178:77 – 85.

Lowe, R. J., Falter, J. L., Kose↵, J. R., Monismith, S. G., and Atkinson,M. J. (2007). Spectral wave flow attenuation within submerged canopies:Implications for wave energy dissipation. Journal of Geophysical Research:Oceans, 112(C5):n/a–n/a. C05018.

Lowe, R. J., Kose↵, J. R., and Monismith, S. G. (2005). Oscillatory flowthrough submerged canopies: 1. velocity structure. Journal of GeophysicalResearch, 110:doi:10.1029/2004JC002788.

Luhar, M., Coutu, S., Infantes, E., Fox, S., and Nepf, H. M. (2010). Wavein-duced velocities inside a model seagrass bed. Journal of Geophysical Re-search, 115:doi:10.1029/2010JC006345.

Mazda, Y., Magi, M., Ikeda, Y., Kurokawa, T., and Asano, T. (2006). Wavereduction in a mangrove forest dominated by sonneratia sp. WetlandsEcology and Management, 14:365 – 378.

27

Page 28: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

Mei, C. C. (1989). The applied dynamics of ocean surface waves. WorldScientific, New Jersey.

Mo↵ett, K. B., Nardin, W., Silvestri, S., Wang, C., and Temmerman, S.(2015). Multiple stable states and catastrophic shifts in coastal wetlands:Progress, challenges, and opportunities in validating theory using remotesensing and other methods. Remote Sens., 7:10184 – 10226.

Mullarney, J. C. and Henderson, S. M. (2010). Wave-forced mo-tion of submerged single stem vegetation. J. Geophys. Res.,115:doi:10.1029/2010JC006448.

Mullarney, J. C., Henderson, S. M., Reyns, J. A. H., Norris, B. K., andBryan, K. R. (2016). Analysis of tidal flow rotation within a wave-exposedmangrove forest. Continental Shelf Research, submitted.

Nardin, W., Woodcock, C. E., and Fagherazzi, S. (2016). Bottom sedimentsa↵ect sonneratia mangrove forests in the prograding mekong delta, viet-nam. Estuarine, Coastal and Shelf Science, 177:60 – 70.

Nicholls, R., Hoozemans, F., and Marchand, M. (1999). Increasing floodrisk and wetland losses due to global sea-level rise: regional and globalanalyses. Global Environmental Change - Human and Policy Dimensions,9:S69 – S87.

Norris, B. K., Mullarney, J. C., Bryan, K. R., and Henderson, S. M. (2016).The e↵ect of pneumatophore density on turbulence: A field study in aSonneratia-dominated mangrove forest, Vietnam. Continental Shelf Re-search, page submitted.

Pujol, D., Casamitjana, X., Serra, T., and Colomer, J. (2013a). Canopy-scaleturbulence under oscillatory flow. Continental Shelf Research, 66:9 – 18.

Pujol, D., Serra, T., Colomer, J., and Casamitjana, X. (2013b). Flow struc-ture in canopy models dominated by progressive waves. Journal of Hydrol-ogy, 486:281 – 292.

Ri↵e, K. C., Henderson, S. M., and Mullarney, J. C. (2011). Wavedissipation by flexible vegetation. Geophysical Research Letters,38:doi:10.1029/2011GL048773.

28

Page 29: Wave-frequency Flows Within a Near-bed Vegetation Canopy · 2017-03-22 · (Weitzman et al., 2015) and random waves, and tested or calibrated against observed wave attenuation (Jadhav

Siikamaki, J., Sanchirico, J. N., and Jardine, S. L. (2012). Global economicpotential for reducing carbon dioxide emissions from mangrove loss. Pro-ceedings of the National Academy of Sciences of the United States of Amer-ica, 109(36):14369–14374.

Sumer, B. and Fredsøe, J. (1997). Hydrodynamics Around Cylindrical Struc-tures. Advanced Series in Applied Physics. World Scientific.

Tanino, Y. and Nepf, H. M. (2008). Laboratory investigation of mean dragin a random array of rigid, emergent cylinders. Journal of Hydraulic En-gineering, 134(1):34–41.

Temmerman, S., Meire, P., Bouma, T., Herman, P., Ysebaert, T., andDe Vriend, H. (2013). Ecosystem-based coastal defence in the face ofglobal change. Nature, 504:79 – 83.

Thu, P. M. and Populus, J. (2007). Status and changes of mangrove forestin Mekong Delta: Case study in Tra Vinh, Vietnam. Estuarine, Coastaland Shelf Science, 71(1-2):98–109.

Tolman, H. (1994). Wind waves and movable-bed bottom friction. Journalof Physical Oceanography, 24:994 – 1009.

Weitzman, J. S., Zeller, R. B., Thomas, F. I. M., and Kose↵, J. R. (2015). Theattenuation of current- and wave-driven flow within submerged multispe-cific vegetative canopies. Limnology and Oceanography, 60(6):1855–1874.

Welch, P. D. (1967). The use of fast Fourier transform for the estimationof power spectra: A method based on time averaging over short, mod-ified periodograms. I.E.E.E. transactions on audio and electroacoustics,15(2):70–73.

Zeller, R. B., Zarama, F. J., Weitzman, J. S., and Kose↵, J. R. (2015). A sim-ple and practical model for combined wavecurrent canopy flows. Journalof Fluid Mechanics, 767:842880.

29