Lim Et Al 2016

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Coastal Engineering 116 (2016) 258–274

Contents lists available at ScienceDirect

Coastal Engineering

journal homepage: www.elsevier.com/locate/coastaleng

An experimental study on near-orthogonal wave–current interaction


over smooth and uniform fixed roughness beds
Kian Yew Lim a,⁎, Ole Secher Madsen b
a
Department of Civil and Environmental Engineering, National University of Singapore, 2 Engineering Drive 2, EW1 03-01A, 117577, Singapore
b
Department of Civil and Environmental Engineering, Ralph M. Parsons Laboratory, 48-317, MIT, Cambridge, MA 02139, USA

a r t i c l e i n f o a b s t r a c t

Article history: An experimental study of periodic waves interacting with near-orthogonal turbulent currents is presented in this
Received 31 December 2014 paper. Mean velocity profiles for current-alone, wave-alone and combined wave–current flows are measured
Received in revised form 3 May 2016 with Acoustic Doppler Velocimeters to resolve the changes in mean flow kinematics due to wave–current inter-
Accepted 11 May 2016
action. In this study, the near-bottom wave orbital velocity is approximately 1.5 times greater than the depth-
Available online 5 August 2016
averaged current velocity, and the log-profile method is used to determine the bottom roughness from the mea-
Keywords:
sured velocity profiles. The primary focus is on 90° wave–current interaction, while selected findings for 60° and
Wave–current interaction 120° wave–current orientations are also presented. In the smooth bed experiment, the interaction is shown to be
Boundary layer flow linear due to the relatively low Reynolds-number flows generated in the facility — a relatively common scenario
Bottom roughness in small-scale laboratory setups that has not received much attention in previous studies. The smooth turbulent
Grant–Madsen model flow equation, with modification of the input shear velocity, is found to accurately predict the mean flow rough-
ness for linear interaction. The bottom roughness is subsequently increased with the introduction of a layer of
uniform 12.5 mm marbles to achieve a more realistic rough turbulent flow regime. The results agree qualitatively
with previous experimental findings that showed a reduction in the near-bottom mean velocity due to a wave-
enhanced (apparent) roughness. However, the Grant–Madsen (GM) model is found to over-estimate the appar-
ent roughness when the angle of wave–current interaction is large, implying a lack of directional sensitivity of
this model under “strong-wave, weak-current” conditions. A tentative explanation of this shortcoming is given
in the paper. However, it is not possible to extend the GM model analysis to the present near-orthogonal
wave–current flows, as results suggest that both 60° and 120° wave–current cases are sufficiently contaminated
by the wave-induced mass transport component in the current direction to invalidate the use of the log-profile
method to resolve the bottom roughness. In addition to the modification of current profiles by waves, the present
study also shows a drastic transformation of wave-induced mass transport profiles by the external turbulent cur-
rents. The veering of mean flow over depth is consistent with the superposition of a wave-induced return flow on
the external current, while additional veering in the near-bottom region may be attributed to turbulence asym-
metry induced by the current component in the direction of the near-bottom wave orbital velocity which is
shown to vary locally by ±10° due to parasitic waves emanating from the current inlet and outlet.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction interaction is restricted to a thin layer of water column close to the sea-
bed, further complicated by the presence of bedforms in many cases.
It is commonly recognized that coastal circulation and sediment Since this interaction is a highly nonlinear process, the combined
transport are significantly influenced by combined waves and current wave–current flow is considerably different from a direct summation
flows. In coastal waters, the direction of wave propagation is predomi- of current-alone and wave-alone kinematics, implying that meaningful
nantly near-orthogonal to currents, and both flows exhibit considerable studies require both currents and waves to be present simultaneously.
differences in time scales. The period, T, of coastal currents is of the Immense progress has been made in the last few decades on the
order of several hours, and the boundary layer thickness, which is simulation of wave–current flows, starting from the simple models de-
pffiffiffi
roughly proportional to T, is of the order of meters. Wind-wave pe- veloped in the 1970s and 1980s [Lundgren (1972), Smith (1977), Grant
riods are usually less than 20 s and suggest that the wave boundary & Madsen (1979), Fredsøe (1984), Christofferson & Jonsson (1985),
layer thickness is only of the order of centimeters. Hence, wave–current Coffey & Nielsen (1986), Myrhaug & Slattelid (1989)], to more recent
numerical models that adopt state-of-the-art turbulence closure
⁎ Corresponding author. methods, e.g. Holmedal et al. (2013) for collinear wave–current flows,
E-mail addresses: dprlky@nus.edu.sg (K.Y. Lim), osm@mit.edu (O.S. Madsen). Davies et al. (1988), Olabarrieta et al. (2010) and Afzal et al. (2015)

http://dx.doi.org/10.1016/j.coastaleng.2016.05.005
0378-3839/© 2016 Elsevier B.V. All rights reserved.
K.Y. Lim, O.S. Madsen / Coastal Engineering 116 (2016) 258–274 259

for arbitrary angles of wave–current flows. Detailed review of this sub- discussed. In addition to the modification of current flows by waves,
ject can be found in Grant & Madsen (1986) and Soulsby et al. (1993). A the present study is also aimed at understanding the changes in wave-
primary strategy of early-day theoretical studies was to adopt simple induced mass transport flows due to wave–current interaction, a phe-
turbulence closure models when solving the Navier–Stokes equations nomenon that has not received much attention in past studies but
for both steady and oscillatory components of the combined flow. The could have significant implications for cross-shore sediment transport
available closure models include the eddy viscosity model, the turbulent since there is a strong dependency of bedload transport on near-bed
kinetic energy closure method and the momentum-deficit integral mean currents.
method (Soulsby et al., 1993). The eddy viscosity model is one of the The organization of this paper is as follows: Section 2 discusses the
simplest to use, and wave–current models based on this scheme are experimental setup and preliminary tests used to establish the mea-
generally integrated into large-scale ocean circulation models to en- surement regions and uncertainties. The smooth bed experiments for
hance computational efficiency of numerical simulation. 90° wave–current interaction are detailed in Section 3. Section 4 pre-
Validation of theoretical models has been performed mostly with sents the 90° flow experiments conducted over a uniform fixed rough-
collinear wave–current experiments (e.g. Kemp & Simons, 1982, 1983; ness bed and compares them with predictions afforded by the GM
Klopman, 1994; Mathisen & Madsen, 1996a, b; Fredsøe et al., 1999), model. Potential mechanisms influencing the veering of near-bottom
though a fair amount of experimental work involving near-orthogonal mean flow are also discussed. Selected results on 60° and 120° wave–
wave–current interaction has also been conducted (e.g. Havinga, current interaction are presented in Section 5, with discussions of the
1992; Simons et al., 1992; Arnskov et al., 1993; van Rijn & Havinga, contamination of log-profile analysis by near-orthogonal waves. The
1995; Musumeci et al., 2006; Madsen et al., 2008; Fernando et al., final section presents conclusions of the study and recommendations
2011). The general conclusion is that when waves are present with cur- for future work.
rents, the near-bottom mean velocity increases (decreases) over
smooth (rough) beds. Besides, collinear wave–current studies have 2. Experiments and methodology
also demonstrated that for currents following (opposing) the waves,
the near-surface mean velocity decreases (increases) with respect to 2.1. Experimental setup
current-alone conditions (Kemp and Simons, 1982, 1983; Klopman,
1994). When waves interact with currents in orthogonal or near- Experiments were performed in the Hydraulic Engineering Labora-
orthogonal orientations, the mean flows would experience changes tory of the Department of Civil & Environmental Engineering, National
not only in terms of magnitude but also in the flow directions due to University of Singapore (NUS). Current-alone, wave-alone and com-
the superposition of wave-induced return flow on the nominal current, bined wave–current flows were generated in a 33 m (L) × 10 m
as well as steady streaming induced by momentum transfer (Longuet- (W) × 0.9 m (D) wave–current basin. A photo of the facility is shown
Higgins, 1953 — henceforth LH53) and turbulence asymmetry in Fig. 1 while the overall layout is presented in Fig. 2. Water depth, h
(Trowbridge & Madsen, 1984) within the wave boundary layer. Numer- was maintained at 0.4 m for all experiments. A section of the basin
ical simulations by Davies et al. (1988 — henceforth DSK88), accounting (19 m × 1.5 m) parallel to the direction of the wave propagation was
only for turbulence asymmetry, and Afzal et al. (2015 — hereafter converted into a reservoir to generate steady current flows in the test
AHM15) who also included mean momentum transfer, have revealed area. The flows exited the basin over an adjustable weir at the outlet
details of this veering of the mean flow for non-collinear wave–current and were re-circulated to the reservoir by two 75HP centrifugal
flows. However, to the authors' knowledge, these details have not been pumps. Current inlets of width 2.5 m were constructed along the wall
supported by experimental evidence. of the reservoir and aligned such that currents would intersect with
The comparison between theoretical models and experimental re- waves at 60°, 90° and 120° when both flows were present. For notation
sults require a precise knowledge of the bed resistance over which the purpose, the current channel that connects Inlet 2 with the outlet is
flows interact, and this is commonly expressed as the equivalent known as the 90° current channel, while those that connect Inlets 1
Nikuradse sand grain roughness (henceforth bottom roughness, kn). and 3 with the outlet are denoted as the 60° and 120° current channel,
For the case of steady rough turbulent flows over a uniform flat sand respectively (refer to Fig. 2). Honeycomb filters consisting of 50 cm-
bed, kn is characterized by the diameter of the sand grains, with numer- long, 5 cm-diameter PVC pipes were installed at the inlets to ensure
ous studies suggesting that kn is of the order 1 to 10 d50 (Nielsen, 1992).
Several notable studies on oscillatory flows over movable beds (Carstens
et al., 1969; Lofquist, 1986) conclude that kn is approximately four times
the ripple height, η, when bedforms are present. However, for oblique
wave–current flows in a basin, the movable bedforms tend to exhibit
greater spatial variability (Madsen, et al., 2008), which results in ambigu-
ity of the value of kn. This problem can be partly overcome by using fixed
2D bedforms, but strong directional-dependency of kn is usually ob-
served for this type of bed configuration due to drastic changes in bed re-
sistance when the angle between mean flow and bedform axis varies
(Barrantes & Madsen, 2000). The above shortcomings may be resolved
by generating wave–current flows over a uniform, fixed 3D bottom con-
figuration, where kn is spatially homogenous as well as directional-
independent. Unfortunately, experimental measurements involving
near-orthogonal waves and currents over this type of bed configuration
are relatively scarce.
The aforementioned gap motivates the present study, which is
aimed at obtaining high-quality experimental data for near-
orthogonal wave–current flows over smooth and uniform fixed rough-
ness beds. The focus is on waves interacting with currents at 90°, and a
large number of near-bed measurement data were collected to facilitate
the determination of bottom roughness with the log-profile analysis. Fig. 1. Bird's eye view of the wave–current basin in the Hydraulic Engineering Laboratory,
Selected findings on 60° and 120° wave–current flows are also NUS.
260 K.Y. Lim, O.S. Madsen / Coastal Engineering 116 (2016) 258–274

Fig. 2. Layout of wave–current basin. Shaded areas in the middle of the basin denote the near-uniform current flow regions (to be discussed in Section 2.3.2).

uniformity and directionality of the inflows. During the experiment, the anticlockwise from the positive x-direction, which is the direction of
flow discharge was constantly monitored with an electromagnetic flow wave propagation (θR = 0°).
meter. Two types of bed conditions were examined in the present study, The presence of current along the direction of wave propagation (for
i.e. a smooth concrete bed and a uniform roughness bed with ceramic oblique wave-current cases) may change the wave properties shown in
marbles of d50 = 12.5 mm covering the concrete bottom. Table 1, but this was found to be negligible. For instance, with Ta =1.4 s
Waves were generated in the basin by a multi-element, piston-type and ωr ¼ ωa −ku, where ωr is the relative radian frequency and ωa is the
wave-maker, each paddle controlled by an electric-servo motor. A 1:8 absolute radian frequency (without current), the relative wavelength,
passive absorbing beach was installed at the end opposite from the pad- Lr, is approximately 2.5 m, which is merely 4% different from the abso-
dles for wave absorption, with reflection coefficient estimated to be less lute wavelength, La (when waves are alone).
than 8% for the waves generated in this study. The wave paddles were
located at least 2 m away from the current channels, ensuring that the
influence of evanescent waves, which are significant within ~ 3 times 2.2. Velocity and free surface elevation measurement
the water depth from the wave-maker, can be safely neglected in the
measurement regions (Madsen, 1970). The wave and current condi- Velocity measurements in the basin were performed with Acoustic
tions generated in the 90° current channel are summarized in Table 1, Doppler Velocimeters (Nortek's Vectrino Plus model). The sampling
and the coordinate system used is: positive-x in the direction of the volume was cylindrical in shape with a diameter of 0.6 cm and length
wave propagation, positive-y in the direction perpendicular to the x- of 0.7 cm. It was approximately 5 cm below the probe tip to minimize
axis in a horizontal plane towards the outlet, and positive-z vertical up- interference of the flow field by the physical body of the probe. The sam-
ward from the theoretical bottom. The theoretical bottom for smooth pling frequency was fixed at 200 Hz in all tests. Fine clay powder was
bed is assumed to be the same as the concrete floor level, and for the used as seeding particles to enhance signal reflection from the flow
uniform roughness case it is taken as 0.3 diameter below the top surface and hence the quality of measurements. Data with an acceleration of
of marbles (see Section 2.4 for further description). Velocity components more than 1 g (981 cm/s2) were classified as spikes and filtered from
along the x-, y- and z-axes are denoted by u, v and w, respectively, while the record. Fourier analysis was then conducted on the remaining data
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
the resultant horizontal mean velocity is denoted by uR (¼ u2 þ v2 Þ. to obtain time-averaged and harmonic velocities.
From an aerial view of the basin, the angle of mean flow, θR, increases With the sampling volume located 5 cm below the probe tip, the up-
permost 5 cm of the water column was out of the measuring range.
When waves were present, the highest measurable point was 5 cm
below the wave trough, since the probes were required to be sub-
Table 1 merged at all time. Separate instruments were normally used for near-
Experimental conditions in the 90° current channel.
surface flow measurements. For instance, Fredsøe et al. (1999) used a
Currents 
Measured depth-averaged velocity, U Rec ¼ Uh

bi-directional micro-propeller to measure the flow between wave crests
υ
(cm/s) and troughs, and a Laser Doppler Anemometer (LDA) to cover the re-
CS: smooth bed 11.0 ± 0.1 55,000 maining part of the water column. Since the present study is concerned
CU: uniform roughness bed 10.6 ± 0.2 53,000 mainly with the near-bottom mean flow kinematics, in particular the
region where the logarithmic velocity profile is expected to prevail, no
Waves T(s) ubm (cm/s) Rew = ubmυAbm Abm
kn further attempt was made to measure the near-surface flow.
WS: smooth bed 1.4 16.9 ± 0.6 7,800 – The free surface elevation was measured by several capacitance-
WU1: uniform roughness bed 1.4 16.6 ± 0.5 7,700 1.24 type wave gauges at a frequency of 50 Hz. Up to three wave gauges
WU2: uniform roughness bed 1.6 17.7 ± 0.6 9,900 1.50
were set-up on tripods close to the wave-maker to monitor the station-
Wave–current arity of incident waves, while an additional wave gauge (alongside
three units of ADVs) was mounted on the transverse carriage (see Fig.
WCS: CS + WS
WCU1: CU + WU1
1) and moved along the x- and y-axis of the wave–current basin for de-
WCU2: CU + WU2 tailed measurement within the current channels. The wave gauge was
fixed at a specific elevation, whereas the ADVs were connected to ad-
T = wave period; ubm = measured near-bottom orbital velocity amplitude
υ = kinematic viscosity = 0.008cm2/s at temperature of 30 °C. justable point gauges to allow changes of probe elevation to capture ve-
Abm = near-bottom particle excursion amplitude = ubmT/2π. locity profiles over depth.
K.Y. Lim, O.S. Madsen / Coastal Engineering 116 (2016) 258–274 261

2.3. Preliminary tests the current channel's centerline. A similar approach was used to identify
the near-uniform flow regions in the 60° and 120° channels. The distri-
2.3.1. Sampling duration bution of measurement stations within the near-uniform flow regions
To establish a sampling duration for velocity measurements, an in the three current channels are shown in Fig. 3. Up to 25 points
hour-long velocity record was collected for a steady current flow condi- were measured along the water column at each station to obtain well-
tion in the basin. The record was then subdivided into smaller time se- defined velocity profiles and ensure that sufficient data points were
ries to determine a duration which gave a mean velocity with an available for log-profile fitting. The upper limit for log-profile fitting
acceptable level of uncertainty, i.e. similar to the approach adopted by was set at 1/3 of the current boundary layer thickness predicted with
Mathisen & Madsen (1996a) and Chanson (2008). The standard devia- Eqs. (1) and (2), while the lower limit was taken as twice the diameter
tions, σ, of the sub-divided time series are tabulated in Table 2. When of the roughness element, or 7 mm from the concrete floor for smooth
the sampling duration is 3 min, σ is approximately 0.3 cm/s, which is bed experiments. Adopting these spatial limits for log-profile fitting typ-
less than 3% of the time-averaged current velocity. By shortening the ically resulted in 6–9 data points for smooth bed cases and 11–13 data
sampling duration to 1 min, σ remains less than 5% of the time- points for rough bed cases. For wave–current flow, the lower limit
averaged velocity for the hour-long measurement record. would be dictated either by the above criterion or the wave boundary
For consistency, a fixed sampling duration was used in current- layer thickness, whichever was larger.
alone, wave-alone and wave–current experiments. The duration should An example of the spatially-averaged velocity profile over a uniform
therefore cover a sufficient number of wave cycles to allow accurate de- roughness bed is presented in Fig. 4a. For clarity, data points are
termination of wave properties. In this study, waves with periods of grouped into vertical bins of size 1 cm when presented in spatially-
1.4 s and 1.6 s were generated. Based on the suggestion of Sleath averaged velocity plots (i.e. points within ±0.5 cm from a mean z posi-
(1987) that no less than 50 wave cycles should be used for phase aver- tion would be averaged and shown as one data point). Since the near-
aging, a 3-min sampling duration, corresponding to more than 100 uniform flow regions were first established based on current-alone
wave cycles, was adopted for all measurement cases in the present flows, the uniformity of these regions for combined wave–current sce-
study. narios was also checked in subsequent tests. It was found that the
near-uniform flow regions were generally valid for both current-alone
2.3.2. Near-uniform flow regions and wave–current flows. For instance, the standard deviations of
The inlet conditions were adjusted such that the near-uniform flow wave–current mean flow over uniform roughness bed in the 90° current
region in each current channel was established close to the center of the channel (Fig. 4b), while obviously greater than their current-alone
wave–current basin. These regions should be sufficiently far from the counterparts, are still of the order 0.5 cm/s and comparable to the tem-
inlet to allow the development of current boundary layers, yet main- poral uncertainty of the measurement.
taining an adequate distance from the outlet to avoid the influence of However, in the experiment for 90° wave–current flow over smooth
the weir. The boundary layer thickness, δc, is predicted from the formu- bed, three up-wave locations in the measurement region (along
las and graphs given by Schlichting (1968), an approximation of which x = −0.25 m in Fig. 3a) were found to have notably smaller mean ve-
was obtained by Madsen (2009) as follows: locities than the other six measurement stations. For this particular
case, the measurement region was “truncated” by omitting the stations
 0:08
υ 0:92
located in the non-uniform section of the region. While the reason for
δc ¼ 0:09 l for smooth bed ð1Þ this deviation is unknown, it is unlikely to have significant implications
U∞
since the remaining number of measurement stations in the “truncated”
0:16 0:84 region is sufficiently large to establish the wave–current flow
δc ¼ 0:13ðkn Þ l for rough bed ð2Þ
characteristics.

where U∞ is the free-stream velocity (approximately equal to the depth-


2.4. Log-profile analysis and bottom roughness
averaged velocity) and l is the distance from the inlet along the current
channel axis (l = y for the 90° case). Based on Eqs. (1) and (2), the
Based on Prandtl's mixing length hypothesis and assumption of a
boundary layer thickness is expected to exceed 10 cm at locations
constant shear stress layer close to the bed, the near-bed mean velocity
more than 3.5 m away from the inlet. Since this should give a satisfacto-
distribution can be described by the logarithmic law:
ry number of measurement points for log-profile fitting, the upstream
and downstream limits of the measurement region along the y-axis  
uc z
were fixed at 3.5 m and 5 m, respectively, from the inlet. In addition, uR ¼ ln ð3Þ
κ z0
preliminary measurements were conducted along the spanwise direc-
tion of the current channel (x-axis) to establish the up-wave and
where u⁎c is the current shear velocity, κ is the von Karman's constant
down-wave limits of the near-uniform flow region. The length of the
(= 0.4), z = z′ + Δ z where z′ = elevation from a fixed datum and
measurement domain along x-axis was progressively narrowed until
Δz = distance between the theoretical bottom and the datum. The log-
an acceptable standard deviation of the order 0.5 cm/s was achieved.
arithmic velocity profile is in principle valid only over a small fraction of
Using this approach, a width of 0.5 m was obtained for the measure-
the bottom boundary layer, with upper limits ranging from 0.1δc
ment region in the 90° current channel, i.e. 0.25 m on either side of
(George, 2007) to 0.3δc (Fredsøe et al., 1999). For rough bed scenarios,
a major difficulty in applying Eq. (3) is the ambiguous location of the
theoretical bottom where the no-slip boundary condition applies.
Grass (1971) proposed an approach to resolve this issue, which involves
Table 2 the plotting of semi-log profiles with different assumptions of Δz and
Standard deviations for datasets subdivided from an hour-long current velocity
identifying the value of Δz that yields the best-fit of Eq. (3) to the exper-
measurement.
imental data. Jackson (1981) concluded that by taking the theoretical
Sampling duration (minutes) Standard deviation (cm/s) bottom as 0.3D below the top surface of the roughness elements, with
10 0.24 D representing the characteristic length (height or diameter) of the
5 0.25 roughness elements, satisfactory results can be obtained for many com-
3 0.33 mon types of rough beds. This level was also found to be a good repre-
1 0.47
sentation of the theoretical bottom by Yuan et al. (2012), who
262 K.Y. Lim, O.S. Madsen / Coastal Engineering 116 (2016) 258–274

Fig. 3. Measurement stations within the near-uniform flow regions for (a) 90° current channel (b) 60° current channel (c) 120° current channel. x′ denotes the local channel axis, with
origin along the centreline of the current channel.

performed their rough bed experiment with a single layer of uniformly- approach is the relatively large uncertainty for z0 when the roughness
arranged marbles identical to those used in the present study. Based on is small and there are few data points available for regression analysis,
these earlier works, an elevation of 0.3D below the top surface of the notably for smooth bed cases where the boundary layer growth is
roughness elements was adopted as the theoretical bottom for the pres- slow. An alternative method of analysis, with ln(z) as the dependent
ent uniform roughness experiment. For the smooth bed experiment, the variable, was performed. Comparison shows that the uncertainty for
theoretical bottom is intuitively the same as the concrete floor level of z0 obtained with the latter approach is considerably smaller than the
the basin. former, with error factors reduced by 2 to 4 times for most locations.
Once Δz was established, Eq. (3) was used to determine the bottom Since z0 is the parameter of interest in this study, the use of ln(z) as
roughness, kn(=30z0). Conventionally, the log-profile fitting of experi- the dependent variable was chosen for smooth bed analysis. The differ-
mental data with Eq. (3) is done with uR as the dependent variable and ence between the two methods is, however, insignificant for the case of
ln(z) as the independent variable. An undesirable outcome of this uniform roughness bed, since the bottom roughness is significantly

Fig. 4. Mean velocity, uR, and one-standard-deviation error bar for (a) current-alone (b) wave–current mean flow over a uniform roughness bed in the 90° current channel. Similar results
were obtained for 60° and 120° current channels. Dashed lines indicate the regions analyzed with log-profile method.
K.Y. Lim, O.S. Madsen / Coastal Engineering 116 (2016) 258–274 263

larger and the boundary layer development is accelerated by the pres- the following expression (Schlichting, 1968):
ence of marbles, resulting in a larger number of data points within the
boundary layer and a better quality of log-profile fitting. υ
kn;smooth ¼ 30z0 ¼ 3:3 ð7Þ
Two filtering schemes were used in this study to eliminate outliers. uc
First, if the shear velocity, u⁎c, determined from log-profile analysis at
a measurement station had a 95% Confidence Interval (C.I.) exceeding which gives kn,smooth = 0.051 cm, i.e. only 15% smaller than the exper-
50% of the mean value, the station would be dropped from further anal- imental kn and well within the experimental confidence interval. Hence,
ysis. The remaining stations that satisfied the first test were then it can be concluded that the current-alone flow experiences a smooth
checked with the Median Absolute Deviation method (Pham-Gia & turbulent roughness when generated over the concrete bed of the basin.
Hung, 2001), which helped to eliminate spatial outliers before the
spatially-averaged results were computed. In most cases, the number 3.2. Wave-alone
of locations rejected by this criterion was at most two (out of the nine
measurement stations available). Wave-alone measurements were performed at measurement sta-
For measurement stations that satisfied both filtering schemes de- tions within each of the near-uniform flow region shown in Fig. 3. In
scribed above, the bottom roughness at these stations were converted order to maintain the same sidewall boundary conditions as wave–cur-
into natural log prior to averaging, i.e.: rent experiments, the current inlet and outlet were left open during
wave-alone experiments. The wave period, T, is 1.4 s and spatially-
1 X averaged ubm is 16.7 ± 1.2 cm/s (or ±7.2%), where the uncertainty de-
lnkn ¼ ln kn ¼ M ð4Þ
n notes one standard deviation of the data. Note that the uncertainty is
greater than that shown in Table 1, as it is computed based on measure-
X 2 ment stations across all the three channels illustrated in Fig. 3. Rew =
ln kn − ln kn
σM ¼2
: ð5Þ ubmAbm/υ is therefore less than 9,000 and the wave-alone boundary
n−1 layer is expected to be laminar based on the regime delineation pro-
posed by Jonsson (1966) and Davies & Villaret (1997).
When the results were converted back to linear scale, the spatially- Fig. 6 shows the spatially-averaged wave-induced mass transport
averaged kn was obtained together with a standard deviation factor, flow profile. The conduction solution proposed by LH53 for the Eulerian
eσM (hereafter referred to as the error factor): wave-induced mass transport flow may be expressed as follows:

8 
< eM  eσ M −upper limit of CI a2 ωk 
kn ¼ eM ð6Þ uE ¼ 2
3 þ kh sinhð2khÞ 3μ h 2 −4μ h þ 1 ð8Þ
: M σM 4 sinh kh
e =e −lower limit of CI  

sinhð2khÞ 3  2
þ3 þ μ h −1
2kh 2
The advantage of this approach is that the lower confidence limit of
kn remains positive at all time, which is desirable as negative values of kn
where a is the wave amplitude, ω is the wave radian frequency, k is the
are physically unrealistic.
wave number, μh = 1 − z/h and the near-bottom velocity (μh = 1) is
3ubm 2 =4c (c = wave celerity = ω/k). Since the LH53 solution
3. Orthogonal (90°) wave–current flow over smooth bed
was computed based on measured ubm, the results are subject to uncer-
tainty in establishing the values of ubm in the basin, e.g. due to the con-
3.1. Current-alone
tamination by wave reflection. This uncertainty can be quantified in
 is approximately 11 cm/s terms of the standard deviation of measured near-bottom orbital veloc-
The depth-averaged current velocity, U,
ity at all stations, which is approximately 7.2% of ubm. The uncertainty of
and measurements were made in the near-uniform flow regions
 E, which is scaled by ubm 2, should therefore be approximately twice the
u
shown in Fig. 3. An example of the velocity profile, plotted in semi-log,
error of ubm, i.e. 14.4%. The theoretical result is plotted in Fig. 6 with the
is presented in Fig. 5 for current-alone in the 90° current channel.
experimental data.
From analysis, the spatially-averaged kn is 0.059 cm with an error factor
It should be noted that the flow generated in the present experiment
of 2.0, while the current shear velocity, u⁎c = 0.52 cm/s ± 9.1%. The
is turbulent, since the Stokes' prediction of a wave-induced return
smooth turbulent bottom roughness, kn , smooth, can be predicted with

Fig. 6. Wave-induced mass transport flow over a smooth bed. Error bars for experimental
Fig. 5. An example of current velocity profile over a smooth bed in the 90° current channel. data represent one standard deviation. Error bars for the theoretical solution are
Points located between the two dashed lines are used for log-profile fitting. associated with ±14.4% of u. Cross: Experimental data; Full Line: LH53 model.
264 K.Y. Lim, O.S. Madsen / Coastal Engineering 116 (2016) 258–274

 ES , in a closed basin (Nielsen, 1992):


current, u the laminar LH53 model. The present experimental results suggest that
the “Klopman effect”, e.g. Huang & Mei (2003), must be included in a the-
gH 2 oretical model of wave-induced mass transport currents for such a model
 ES ¼−
u ð9Þ
8ch to be valid over the entire water depth.

where H is the wave height (~10 cm), c is the wave celerity (=1.7 m/s) 3.3. Wave–current interaction
and h is the water depth (=0.4 m), suggests a simple return current of
1.8 cm/s for our experimental conditions, with a corresponding Reh ¼ The current and wave flows described in Sections 3.1 and 3.2 were
 ES h=υ of almost 9,000. On the other hand, the conduction solution pro-
u superimposed at an angle of 90° for case WCS, defined in Table 1. The
posed by LH53 is, strictly speaking, applicable only to laminar waves with measured ubm during wave–current interaction was 16.4 cm/s, which
a/δw bb 1 (where δw is the wave boundary layer thickness). Nonetheless, is well within the uncertainty of ubm obtained for the wave-alone case.
the following reasons help justify the use of LH53 model for comparison Fig. 7a depicts the current-alone and wave–current mean velocity pro-
with the present experimental results. Since the distance between the files, and clearly demonstrates that the introduction of waves on an or-
measurement region and the passive absorbing beach was more than thogonal smooth turbulent current leads to an increase in the near-
10 m, the mass transport current boundary layer predicted by Eq. (1) bottom mean velocity. Correspondingly, the near-surface flow de-
with U∞ = u  ES = 1.8 cm/s is expected to be more than 40 cm in the mea- creases in order to maintain a constant discharge along the current
surement regions. Besides, as measurements were made at least 30 min channel. The increase in near-bottom mean flow agrees qualitatively
after start of wavemaker, our setup and procedures suggest that the vor- with collinear wave–current experiments over smooth beds, e.g.
ticity generated at the bed had diffused over the entire depth at the time Kemp & Simons (1982, 1983). Musumeci et al. (2006) conducted an or-
and locations of measurement. Thus, treating the turbulent mass trans- thogonal wave–current experiment over a fixed bed with relatively
port current as “pseudo-laminar” supports our adoption of the LH53 con- small roughness (d50 = 0.24 mm) and also observed a similar increase
duction solution for comparison with our data. in the near-bottom mean flow upon superposition of waves.
As observed in Fig. 6, the measured wave-induced mass transport flow In addition to the changes in mean velocity, the superposition of
exhibits a near-linear variation over depth and a streaming velocity near waves also modifies the direction of mean flow, θR, over depth. Initially
the smooth bed of about 0.7 cm/s, which is 40% less than the theoretical aligned at 90° for current-alone, θR increases by almost 10° after the in-
result of LH53. As a consequence of LH53's surface boundary condition troduction of waves (Fig. 7b). For a depth-averaged nominal current ve-
for the mean flow, which amounts to imposing a mean shear stress in locity of 11 cm/s, θR is expected to veer by about 9° from its initial 90°
the direction of wave propagation, the LH53 model predicts a reduction orientation based on Stokes theory (Eq. (9)), which is in excellent
of the offshore-directed mean flow as the free surface is approached, i.e. agreement with our experimental observation. The mean velocities in
for z N ~0.6h in our experiment, a trend not exhibited in our measure- the wave direction, u, for current-alone, wave-alone and combined
ments. However, our measurements extended only to ~0.2h below the wave–current flows are presented in Fig. 8. When both current and
free surface, so it cannot be precluded that the mean flow in the offshore waves are present simultaneously, the vertical distribution of mass
direction is reduced in the immediate vicinity of the free surface when the transport flow changes drastically from a near-linearly increasing veloc-
mean flow is turbulent. Similar observations were made by the authors ity profile (for wave-alone) to a near-constant velocity distribution (for
(Lim, 2013) in a flume experiment (h = 0.5 m) for wave periods ranging wave–current flow). This observation agrees qualitatively with the re-
from 1.0 s to 2.6 s (0.57 b kh b 2.10), and also in previous studies, e.g. sults obtained by Musumeci et al. (2006) over fixed roughness elements
Nadaoka & Kondoh (1982), Klopman (1994) and Scandura & Foti of diameter 0.24 mm and 30 mm, and their study with Rew ~ 1,000–
(2011). These observations of a continuous linear decrease of the mean 12,000 showed that changes in u are more significant for small rough-
flow up to ~ 0.2h below the free surface may, at least partially, be ex- ness configuration (Abm/kn ~ 20–90) than large roughness configuration
plained by the observations for collinear wave–current interaction (Abm/kn ~ 0.2–0.7). It appears from Fig. 8 that the u velocity distribution
(Klopman, 1994), where the magnitude of the near-surface current is for wave–current flow can be simply modeled by a uniform return cur-
shown to increase in the presence of opposing waves. For the case of  ES given by Eq. (9), instead of by the more complex LH53 solution,
rent, u
waves interacting with a turbulent wave-induced return current, the except in the near-bottom region where a boundary layer is developing.
flow orientation is exactly the same as waves opposing a current, but While this may be a result of the enhanced mixing within the potential
the near-surface characteristic described above is not accounted for by flow region for the waves by the external turbulent current, a more

Fig. 7. (a) Mean velocity, uR (b) Angle of mean flow, θR, for current-alone and combined wave–current flows over a smooth bed. Full circle: current-alone; Hollow square: wave–current.
K.Y. Lim, O.S. Madsen / Coastal Engineering 116 (2016) 258–274 265

and combined wave–current flow (u⁎c = 0.49 cm/s ± 7.9%). Without


nonlinear enhancement of shear stress, the maximum combined
wave–current shear velocity, u⁎m, can be computed directly using the
results from current-alone and wave-alone:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2 2
um 2 ¼ ðuwm 2 þ uc 2 j cosϕcw jÞ þ ðuc 2 sinϕcw Þ ð12Þ

where u⁎wm is the maximum wave-alone shear velocity and ϕcw is the
angle of wave–current flow, approximately 100° in this case. The max-
imum bottom shear stress for laminar waves can be obtained with the
following expression (Schlichting, 1968):
pffiffiffiffiffiffiffi 1=2
uwm ¼ υωubm ð13Þ

Fig. 8. Mean velocity profiles in the wave direction, u, for current-alone, wave–alone and which gives u⁎wm =1.77 cm/s for the present experimental conditions.
combined wave–current flows over a smooth bed. Dashed line denotes the Stokes
return current (− 1.8 cm/s). Full circle: current-alone; cross: wave-alone; hollow
To estimate the smooth turbulent roughness for a mean flow in the
square: wave–current. presence of waves, which is denoted by kn,cw, the maximum combined
wave–current shear velocity, u⁎m, should be adopted as the input shear
velocity in Eq. (7) instead of current shear velocity, u⁎c:
likely explanation for the drastic change in the u-velocity distribution
over depth from near-linear for wave-alone to near-constant for com- υ
kn;cw ¼ 3:3 : ð14Þ
bined waves and current is obtained by considering the mass transport um
current established in the down-wave region from the nominal current
channel. Since only waves are present here, the mass transport velocity The bottom roughness is expected to decrease by a factor u⁎m/u⁎c of
profile would be near-linear. However, when this pure wave-induced 3.4 from current-alone to wave–current condition, based on Eqs.
mean flow encounters the external current of ~11 cm/s upon entering (7) and (14). The experimental bottom roughness was found to be
the nominal current channel, the established near-linear mass transport 0.059 cm (error factor of 2.0) for current-alone, and for wave–current
current will be advected in the direction of the combined mean flow flows a value of 0.015 cm (error factor of 2.2) was obtained. This trans-
which is approximately at an angle of θR = 100°. Thus, even the water lates into a reduction factor of 3.9 and agrees closely with the theoretical
column entering the nominal current channel at y = 0 would not factor. The relatively large error factor is not surprising given the fact
reach a location further than ytan(10°) into the nominal current chan- that the value of kn , cw is small in this case and any minor changes in
nel. With the location of measurement stations at y ~ 4.3 m this distance slope of the log-profile fit will result in huge variations on the log-axis.
would be approximately 0.75 m and the characteristic, i.e. the near- The overall result suggests that the use of Eq. (14), with u⁎m replacing
linear mass transport current, carried by this water column would u⁎c, is fairly accurate in predicting the bottom roughness when laminar
never reach the measurement locations, which are at least 1 m into waves are present simultaneously with an orthogonal current over a
the nominal current channel. This simple conceptual model suggests smooth bed, provided δw b δvc. Similar conclusion was also obtained by
that water columns passing through the measurement region originate Yuan & Madsen (2014) on the validity of Eq. (7) for collinear wave–cur-
from the current inlet and should exhibit the features of a developing rent flows, though their oscillating water tunnel experiment (with
boundary layer originating from y = 0. This conclusion is supported Rew N 400,000 and wave boundary layer certainly in smooth turbulent
by the mean velocity profiles shown in Figs. 7a and 8 which both sug- regime) suggests that the average wave–current shear velocity should
gest a boundary layer thickness of about 10 cm, in excellent agreement be used as the scaling shear velocity in Eq. (7) for their flow conditions.
with the value obtained from Eq. (1). The above analysis should be a step forward in providing a means to
resolve wave–current interaction over smooth beds. However, this type
3.4. Analysis of smooth bed experiment of flow condition is relatively rare in the field. In order to achieve a more
realistic case of rough turbulent wave–current interaction, the bottom
Since the wave Reynolds number, Rew, over a smooth bed is approx- roughness was increased by covering the smooth bed with a layer of ce-
imately 7,800 in the present study, the wave boundary layer is laminar ramic marbles.
when waves are alone. The laminar wave boundary layer thickness, δw,
is given by Sleath (1987) as follows: 4. Orthogonal (90°) wave–current flow over uniform roughness bed
rffiffiffiffiffiffi
2υ 4.1. Current-alone
δw ¼ 3 ð10Þ
ω
For uniform roughness experiments, the bed was covered with a
which gives δw = 1.8 mm for a 1.4 s wave period. Meanwhile, for uniform, single layer ceramic marbles with d50 = 12.5 mm throughout
smooth turbulent current flow, the viscous sub-layer of the current is the entire basin (Fig. 9). An example of the velocity profile plotted in
approximated by: semi-log scale is shown in Fig. 10. While Eq. (1) is shown to be accurate
in estimating the boundary layer thickness for flows over smooth beds,
υ it appears that the prediction obtained with Eq. (2) for rough beds is less
δvc ¼ 11:6 ð11Þ
uc satisfactory. The boundary layer thickness, obtained from Eq. (2), is ap-
proximately 24 cm, while Fig. 10 shows a thickness of less than 15 cm.
which gives δvc = 1.9 mm for the experimentally determined value of Hence, Eq. (2) over-predicts the boundary layer thickness for uniform
u⁎c = 0.52 cm/s. As δw ≈ δvc for the present experimental conditions, rough beds by more than 60%. Despite this inaccuracy, the results
the laminar wave boundary layer is embedded within the viscous sub- from log-profile analysis are unlikely to be compromised since the anal-
layer of the current flow. This suggests the interaction between the ysis utilizes measurement points within 1/3 of the boundary layer thick-
two flows is purely linear, which is further supported by similar mean ness predicted with Eq. (2). From Fig. 10, it can be seen that this region is
shear velocities obtained for current-alone (u⁎c = 0.52 cm/s ± 9.1%) well-described by the log-law.
266 K.Y. Lim, O.S. Madsen / Coastal Engineering 116 (2016) 258–274

Fig. 9. Uniform, single layer ceramic marbles, held in place by 2 m × 2 m grids formed
using 4 mm-diameter aluminum rods (inset: a close-up view of the marbles). Fig. 11. Wave-induced mass transport velocity over a uniform roughness bed (WU2).
Error bars of the experimental data represent one standard deviation. Error bars for the
theoretical solution are associated with ±11.2% of u. Cross: experimental data; full line:
From the analysis, current-alone kn was determined to be 3.0 cm LH53 model.
with an error factor of 1.5, and u⁎c = 0.92 cm/s ± 9.1%. The flow is
fully rough turbulent since Re⁎ = knu⁎c/υ = 345, and the value of kn is ap- smooth bed, i.e. about 30% of the value predicted by the LH53 solution.
proximately 2.4d50. A comparison was made with a 2D flume experi- The reduction in streaming velocity with increasing bottom roughness
ment using the same marbles as roughness elements. The water depth is in qualitative agreement with the theoretical analysis of Trowbridge
in that experiment was 0.4 m. Two discharges were generated and the and Madsen (1984), who predicted this trend based on a time-varying
theoretical depth-averaged velocities were 10 cm/s and 16.7 cm/s, re- eddy viscosity model. Davies & Villaret (1999) also predicted similar
spectively. Velocity profiles were measured at a distance of 7.5 m from trend for flows over very rough and rippled bottom (Abm/kn b 1.5). Not-
the inlet, and the values of kn obtained were 3.0 cm ( U  = 10 cm/s) withstanding the difference in the near-bed streaming velocity, the var-
and 2.7 cm (U  = 16.7 cm/s), with error factors of 1.2 for both cases. iation of mean flow with elevation is similar for both smooth and rough
These results are in good agreement with the current-alone roughness bed experiments, i.e. a near-linearly increasing velocity in the offshore
obtained in the wave–current basin. Furthermore, the current-alone kn direction as the water surface is approached. This trend was also ob-
also agrees with the wave-alone kn obtained using a wave flume located served in a separate flume experiment with the same marble bed for
in the same laboratory (Lim et al. 2012). In the wave experiment, waves kh = 0.57–2.10 (Lim, 2013). As discussed in Section 3.2, it appears nec-
of 1.6 s period and 10 cm height were generated over smooth and essary to incorporate the near-surface flow characteristic observed by
marble-covered beds, and kn was derived based on the energy dissipa- Klopman (1994) in modeling the wave-induced mass transport flows
tion method. The analysis shows that wave-alone kn over the marble- to achieve a better agreement with experimental observations.
covered bed is 2.7 cm (error factor of 1.3), supporting the conjecture
that a single roughness length scale can be used for both currents and 4.3. Wave–current interaction
waves over uniform 3D bed roughness.
In contrast to the smooth bed experiment, a reduction in near-
bottom mean velocity was observed over the uniform roughness bed
4.2. Wave-alone upon superposition of waves (Fig. 12a). The decrease, which is attribut-
ed to the enhanced turbulence within the wave boundary layer, is in
Similar to the current-alone flow, the wave boundary layer is in the qualitative agreement with existing nonlinear wave–current boundary
fully rough turbulent regime with Rew ~ 10,000 and Abm/kn ~ 1.5 layer models (e.g. the Grant-Madsen (GM) model), and has been ob-
(Jonsson, 1966; Davies & Villaret, 1997). Results for case WU2 is pre- served in numerous collinear (e.g. Kemp & Simons, 1982; Mathisen &
sented in Fig. 11, as measurement data (over depth) for case WU1 are Madsen, 1996b; Fredsøe et al., 1999) and orthogonal (e.g. Musumeci
not available. With the increase in bottom roughness, the near-bed et al., 2006; Madsen et al., 2008; Fernando et al., 2011) wave–current
streaming flow is smaller than the streaming velocity observed over a experiments. The veering of θR is slightly larger for the rough bed exper-
iment - up to 14° from the current-alone direction (Fig. 12b). As in the
smooth bed experiment, such veering can be attributed to Stokes' re-
turn current. The veering in the near-bottom region (z b 10 cm) is asso-
ciated with turbulence asymmetry and mean momentum transfer
(DSK88 and AHM15) and will be discussed further in Section 4.4.
The mean velocity profile in the wave direction, u, was found to be
similar in trend to the smooth bed experiment, where the near-
linearly increasing profile for wave-alone is transformed into a near-
homogenous velocity distribution upon the addition of an external cur-
rent (Fig. 13). As discussed in Section 3.3, the mass transport flow ob-
served in the measurement region is expected to have developed from
the current inlet, since water columns that enter from outside the cur-
rent channel would propagate at a resultant flow direction, θR ~ 105°
and would miss the near-uniform flow region. Again, this is supported
Fig. 10. An example of current velocity profile over a uniform roughness bed in the 90°
by the observed boundary layer thickness for the mass transport flow,
current channel. Points located between the two dashed lines are used for log-profile which is approximately the same as the external current flow which
fitting. originates from the current inlet.
K.Y. Lim, O.S. Madsen / Coastal Engineering 116 (2016) 258–274 267

Fig. 12. Mean velocity, uR, and angle of mean flow, θR, for current-alone and 90° wave–current flows (WCU1) over a uniform roughness bed. Full circle: current-alone; hollow square:
wave–current.

The observed reduction in the near-bottom mean velocity can be ex- of kn ,cw by the GM model is approximately 45% for WCU1 and 67% for
plained by the presence of two layers of log profiles in combined wave– WCU2.
current flows, one within the wave boundary layer and another outside At least part of the observed mismatch between kn , cw and current-
the wave boundary layer but within the current boundary layer. Since alone kn may be attributed to the insensitivity of the GM model to ϕcw,
the wave boundary layer is in the range of 1–2 cm in this study and in particular when the waves are strongly dominating the currents. In
smaller than 2d50, the use of 2d50 as the lower limit in log-profile anal- the GM model, the formulation for the combined wave–current shear
ysis is valid. The log-profile fitting to the data, which are outside the stress is expressed as follows (refer to Appendix A for details):
wave boundary layer, is expected to yield the apparent roughness, kna.

On the other hand, the physical bottom roughness, kn , cw, is required τm ¼ C μ τwm ¼ ð1 þ μ Þτwm for ϕcw ¼ 0 ð15Þ
for comparison with current-alone kn as a means of model validation.
To obtain kn,cw, the following procedure was adopted. First, kna and u⁎c qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi


were obtained from log-profile fitting of the measured velocity data. τm ¼ C μ τwm ¼ ð1 þ μ 2 Þτwm for ϕcw ¼ 90 ð16Þ
An iteration was then performed to obtain kn , cw which, when used as
input to the GM model, yielded a kna value that matched the experimen- where μ is the ratio of the current shear stress to the maximum wave
tal kna. The details of the computational procedures are elaborated in shear stress (τc/τwm). For strong wave conditions where μ approaches
Appendix A. In this study, kna was found to be 7.1 cm (error factor of zero, τm is virtually the same as τwm, indicating that the result is insen-
1.3) and 5.5 cm (error factor of 1.8) for WCU1 (T = 1.4 s) and WCU2 sitive to the angle of wave–current interaction. Additional evidence
(T = 1.6 s), respectively. The analysis shows that kna predicted by the supporting the above argument of a “directional-insensitive” GM
GM model matches the experimental kna when kn , cw = 1.7 cm for model comes from the numerical simulation of DSK88, where ubm =
WCU1 (error factor of 1.4) and kn , cw = 1.0 cm for WCU2 (error factor 100 cm/s, kn , cw = kn = 15 cm, Rew = 1.6 × 106, Abm/kn = 8.5 and
of 2.3). Both results are considerably smaller than the current-alone u⁎c = 5.8 cm/s. For these specifications and ϕcw = 0, the GM model pre-
and wave-alone bottom roughness over the same bed configuration, dicts an apparent roughness kna = 107 cm which is in good agreement
which are in the range of 2.7–3.0 cm. The amount of under-prediction with the value obtained from DSK88's Fig. 12 (~115 cm). Since the esti-
mation of DSK88's kna was done graphically, an uncertainty of ±5 cm is
anticipated. When ϕcw = π/2 is chosen, the GM model predicts kna =
92 cm which is only slightly smaller than the value for ϕcw = 0 and con-
siderably larger than kna obtained from DSK88's Fig. 12 (~63 cm). These
results support the conclusion that the GM model yields acceptable re-
sults for collinear flows, but over-estimates the wave effect on currents
for orthogonal flows.
To improve the original GM model's directional sensitivity, consider
the case of orthogonal wave–current interaction for which the bottom
shear stress, τy, in the current direction is given by:

τy =ρ ¼ −v0 w0 ð17Þ

where v′ and w′ denote the turbulent velocity fluctuations in the y (cur-


rent) and z directions, respectively. Since the waves are in the x-
direction their contribution to turbulent fluctuations in the spanwise di-
rection (y), v0 w , is expected to be smaller than in the streamwise direc-
Fig. 13. Comparison of current-alone, wave-alone and 90° wave–current (WCU1) mean
tion (x), u0 w , i.e.
flows along the wave direction, u, over a uniform roughness bed. Full circle: current-
alone; cross: wave-alone; hollow square: wave–current. Dashed line denotes the Stokes'
return current. v0 w ¼ γ u0 w ð18Þ
268 K.Y. Lim, O.S. Madsen / Coastal Engineering 116 (2016) 258–274

where the reduction factor, γ, may be roughly estimated from the tur- 4.4.1. Parasitic waves
bulence intensity measurements by Klebanoff (1955) as presented in When waves propagate across the current inlet, they induce a differ-
Schlichting (1968, Fig. 18.5), to be: ence in water level (pressure) between the basin and reservoir sides of
the honeycomb filters. The pressure difference fluctuates with wave pe-
!0:5 riod and hence generates an oscillatory flow with the same frequency as
v0 w 2
γ ¼ 0:76 ð19Þ incident waves through the honeycomb filters. Similar oscillatory flow
u0 w 2 is also expected to be generated at the current outlet. Despite the pres-
ence of the honeycomb filters, the nature of these wave disturbances
Since u0 w is scaled by u⁎wm = (fw/2)0.5 ubm, it follows from Eq. (18) that would be akin to the diffracted waves transmitted into the basin (the
orthogonal wave–current interaction may be considered approximately shadow zone) by a train of waves propagating along the outside of the
equivalent to the collinear interaction of a current and an “imaginary” basin walls encountering an open gap of length equal to the width
wave specified by its near-bottom velocity, vbm = γubm. B = 2.5 m of the inlet or outlet openings. These diffracted waves
When the DSK88 orthogonal wave–current case is analyzed in the would emanate from the openings as approximately radial waves
manner suggested above, the GM model predicts an apparent rough- with amplitudes that vary along the crests. For a gap width of B =
ness of 64 cm, in close agreement with DSK88's value of 63 cm. Applying 2.5 m, which happens to coincide with the length, L, of the 1.4 s
this directional modification of the original GM model to the present ex- waves, the spatial wave amplitude variation, i.e. the diffraction coeffi-
perimental results, a value of kn , cw = 2.5 cm was obtained for case cient pattern for B/L = 1, can be found in Johnson (1953). It shows
WCU1, i.e. cutting the difference between kn , cw and kn for current that the wave amplitude decreases with distance from the opening
alone more than in half (18% versus 45%). However, the agreement ob- and is larger in the down-wave than in the up-wave direction. Given,
tained for case WCU2 is still unsatisfactory, although it reduces the dif- with reference to Fig. 3, the rather large distance from the current
ference between kn,cw and kn from 67% to 49%. inlet (4.3 m) and outlet (3.5 m) the radial parasitic waves would be
While the modification of the original GM model has improved its nearly plane progressive waves of constant amplitude as they travel
directional sensitivity, it should be recognized that the suggested mod- across the small area (0.5 m by 0.8 m) of the measurement region, i.e.
ification is limited to orthogonal wave–current flows. Thus, additional the parasitic waves would show up as a partially standing wave in the
theoretical developments are called for to generalize the directional current channel direction that would introduce a non-zero wave veloc-
modification of the GM model to arbitrary angles between waves and ity in the y-direction, ve = vm cos (ωt), where vm would be zero in the ab-
currents. In this respect, it is interesting to note that the reduction factor sence of parasitic waves. Further evidence of the existence of the type of
γ = 0.76, here obtained by considering turbulence anisotropy which is parasitic waves identified in the present experimental set-up and their
not included in neither DSK88's nor AHM15's numerical models, is near- approximate description by the diffraction analogy can be found in
ly equivalent to assuming that the eddy viscosity affecting the orthogo- Fig. 16a of Musumeci et al. (2006), which shows parasitic wave velocity
nal current is scaled by the average (u⁎ave) rather than the maximum amplitudes to decrease from vm ~ 30% to ~ 10% of um when increasing
wave shear velocity (u⁎wm), which would result in a reduction factor the distance from the outlet from 75 cm to 200 cm.
of (2/π)0.5 = 0.80. Thus, to preserve the GM model's ability to predict For notation purpose, the parasitic waves generated at the inlet and
collinear wave–current interaction, it may be necessary to consider tur- outlet are denoted by vi and vo, respectively. A subscript “bm” is used to
bulence anisotropy in the development of an improved theoretical denote the near-bottom velocity amplitudes of the parasitic waves, i.e.
model for wave–current interaction at arbitrary angles. vi , bm and vo , bm. As anticipated from the diffraction analogy described
above, analysis of the pattern of experimentally obtained values of vbm
4.4. Veering of near-bottom mean flow within the measurement region does indeed reflect the pattern associ-
ated with a partially standing wave, and resolving this pattern into its
As shown in the previous sections, the observed mean flows in wave components gives a slightly larger vo , bm ~ 1.85 cm/s than
wave–current cases exhibit a tendency to veer towards the wavemaker vi ,bm ~ 1.12 cm/s as anticipated from the diffraction analogy. Thus, at a
(increase of θR) by 10°–15° over depth, which is consistent with the su- velocity antinode the parasitic wave velocity amplitude is estimated to
perposition of a wave-induced return current on the nominal current be vbm = vi ,bm + vo,bm ~ 3.0 cm/s, i.e. roughly 20% of the velocity ampli-
flow. However, the veering of near-bottom mean flow appears different tude of the “incident” waves propagating in the x-direction, suggesting
from that in the upper column, which may be attributed to other factors that the parasitic waves are a non-negligible source of disturbance in
such as turbulence asymmetry and momentum transfer resulting from the wave basin.
the presence of currents along the wave axis (DSK88 and AHM15). For
co-directional waves and currents, the wave velocity in the half-cycle
following the current is enhanced compared to the half-cycle opposing
the current. This induces an asymmetry in turbulence intensity and gen-
erates a mean stress that drives a current opposite the direction of wave
advance. For waves and currents interacting at oblique angles, this leads
to a veering of near-bottom mean flow towards the axis perpendicular
to the waves (e.g. Fig. 13 of DSK88). The veering trend gradually dissi-
pates with distance from the bed but remains detectable along the
water column even if one is unable to measure within the wave bound-
ary layer.
The above mechanism should clearly manifest for the 60° and 120°
flow cases of the present study (see Section 5). For the 90° case, the
mean flow is generally shifted by 10°-15° (θR ~ 100°–105°) due to a
wave-induced return current, and the near-bottom flow should exhibit
a tendency to veer towards 90° as the bed is approached. However, an
opposite trend is observed in Fig. 12b. With further analysis of the mea-
surement data, it was found that disturbances generated at the inlet and Fig. 14. Changes in the effective wave direction from x to x'', due to parasitic waves being
outlet openings (referred to as “parasitic waves”) may have contributed (a) in-phase and (b) out-of-phase with the incident wave motion. Correspondingly, the
to the observed veering behavior. axis perpendicular to the wave motion changes from y to y' '.
K.Y. Lim, O.S. Madsen / Coastal Engineering 116 (2016) 258–274 269

the near-bottom mean flow to veer towards 90° (Fig. 16b). This is in
qualitative agreement with the conjecture made above, that the near-
bottom veering of mean flow may be locally influenced by the potential
±10° variation of effective wave direction arising from the presence of
parasitic waves.
It should be noted that the conditions leading to the generation of
parasitic waves are inherent in experiments where currents are intro-
duced and removed through sidewall openings. Since these distur-
bances tend to dissipate with distance from the sidewalls, their effects
are likely to be negligible only if the basin width is significantly greater
than the parasitic wave length, a condition clearly not satisfied in the
present setup. Hence, future studies should take into account these dis-
turbances in experimental design and data analysis.

5. Wave–current interaction at 60° and 120° over uniform roughness


bed
Fig. 15. Angle of mean flow, θR, for 90° wave–current flow over a uniform roughness bed,
with dashed lines indicating the variability across the measurement region. 5.1. Mean velocity profiles

When the parasitic waves interact with incident waves, the resultant When waves are superimposed on 60° and 120° current flows, the
wave motion would deviate from the x-direction. For instance, when in- mean velocity profiles, as in the 90° case, exhibit a logarithmic trend
cident and parasitic waves are in-phase, the effective wave direction in the near-bottom region (Fig. 17). The reduction of near-bottom
would rotate anti-clockwise from the x-axis (Fig. 14a). On the other mean velocity after introduction of waves is also consistent with the
hand, when incident and parasitic waves are out-of-phase, the effective GM model. However, the mean velocity decreases substantially for the
wave direction would rotate clockwise from the x-axis (Fig. 14b). Corre- 60° current near the water surface, while it remains virtually unchanged
spondingly, the axis normal to the wave motion would veer by the same for the 120° current. The overall discharge that flowed through the
amount from the original y-axis. Based on the measured strength of ubm near-uniform flow region was smaller compared to the current-alone
and vbm, the veering is expected to be as large as 10°. Assuming a perfect experiment (despite no variation in flow discharge), which was likely
90° mean flow (without the presence of wave-induced mass transport caused by an increase in the lateral spreading of flow when waves
flow), the near-bottom mean flow direction may veer towards 100° at were present.
locations where incident and parasitic waves are in-phase, and towards Current-alone θR in the 60° and 120° channels are shown in Fig. 18. It
80° when incident and parasitic waves are out-of-phase. can be seen that the angles of mean flow in the near-uniform flow re-
Given the continuous change in phase difference between incident gions are close to the desired flow directions, underlining the capability
and parasitic waves throughout the current channel, the effective of the present setup to ensure the directionality of the current inflow.
wave direction (and the axis perpendicular to it) is expected to be With the superposition of waves, θR increases by 10° to 20° which, as
location-dependent. Hence, it is not surprising to observe a considerable discussed in Sections 3.3 and 4.3, can be attributed to Stokes' return
variation of flow veering across the measurement region (Fig. 15). How- flow. The current component in the direction of wave propagation, u,
ever, at some local measurement stations, the veering of flow appears to is presented in Fig. 19. For the 60° case, the strength of u decreases in
correlate with the effective wave direction. As shown in Fig. 16a, when the near-surface region, while the contrary is true for the 120° case.
both incident and parasitic waves are nearly in-phase (effective wave This shows that when a near-orthogonal current flow is resolved into
direction N 0°), the near-bottom mean flow appears to veer towards the direction of wave propagation, the near-surface flow characteristics
110°. In contrast, when incident and parasitic waves are nearly out-of- are similar to co-directional wave–current flows (Klopman, 1994), i.e.
phase (effective wave direction b 0°), there seems to be a tendency for reduction (increase) of the near-surface mean flow when the current

Fig. 16. Local angle of mean flow over a uniform roughness bed at (a) location where incident and parasitic waves are nearly in-phase (b) location where incident and parasitic waves are
nearly out-of-phase.
270 K.Y. Lim, O.S. Madsen / Coastal Engineering 116 (2016) 258–274

Fig. 17. Mean velocity, uR, profiles for current-alone and combined wave–current flows at interaction angles of (a) 60° (b) 120° over a uniform roughness bed. Wave period, T = 1.4 s. Full
circle: current-alone; hollow square: wave–current.

is following (opposing) waves, though the flows in the present study, that could potentially modify the effective wave direction is the current-
unlike Klopman's study, are not fully developed. induced wave refraction, but this was found to be insignificant in the
Parasitic waves are also expected to be present within the measure- present study (i.e. change of less than 2°).
ment regions for the 60° and 120° experiments. The origin and nature of
these parasitic waves are similar to those observed and discussed for the 5.2. Contamination of log-profile analysis
90° case, i.e. over the limited extent of the measurement regions the
parasitic waves will be approximately plane waves propagating in the The spatially-averaged kna, u⁎c and kn , cw for the 60° and 120° cases
direction of the nominal current channel, one coming from the inlet are shown in Table 3. The values of kn,cw were computed in accordance
and the other from the outlet, leading to a partially standing wave pat- with the procedures described in Section 4.3 and Appendix A. For com-
tern. Wave velocity measurements indeed verify this conjecture and pleteness, the results obtained for 90° wave–current flows are also pre-
show that the parasitic wave coming from the inlet is larger (smaller) sented. When the angle between waves and current deviates from 90°,
than the one coming from the outlet for the 60° (120°) case, as expected kna should intuitively be larger since the wave-induced turbulence is in-
from the diffraction analogy presented in Section 4.4. The parasitic wave creasingly aligned with the current flow direction. However, the exper-
velocities in the y-direction cause the effective direction of the local imental outcome is different from this anticipation, with the 60° wave–
wave motion to vary ±10°, just as was the case for the 90° experiments. current interaction yielding the smallest kna (~3.5 cm) among the three
However, in contrast to the 90° case, for which the ±10° variation in ef- angle cases. When the bottom roughness is resolved using the GM
fective wave direction potentially could reverse the veering direction, model, kn ,cw is only 0.5 cm for the 60° case, which is more than a factor
the effect of parasitic waves is not as dramatic for the 60° and 120° ex- of two smaller than the value obtained for the 90° case. On the other
periments since this local variation in effective wave direction is smaller hand, kn , cw for the 120° case (1.4–1.7 cm) is larger than the 90° case,
than the 30° angle between the current and the direction of the normal though only marginally so for 1.4 s waves.
to the incident waves. Indeed, both 60° and 120° flows show a tendency The above observation may be attributed to the contamination of
to veer towards the y-axis (90°) as the bottom is approached (Fig. 18), log-profile analysis by wave-induced mass transport flow. A conceptual
though the trend is somewhat weaker for the 120° case. A second factor explanation of the contamination of the 60° current profile by the wave-

Fig. 18. Angles of mean flow, θR, for current-alone and combined wave–current flows at interaction angles of (a) 60° (b) 120° over a uniform roughness bed. Wave period, T = 1.4 s. Full
circle: current-alone; hollow square: wave–current.
K.Y. Lim, O.S. Madsen / Coastal Engineering 116 (2016) 258–274 271

Fig. 19. Velocity component in the wave direction, u, for current-alone and combined wave–current flows at interaction angles of (a) 60° (b) 120° over a uniform roughness bed. T = 1.4 s.
Full circle: current-alone; hollow square: wave–current.

induced mass transport flow is presented in Fig. 20. First, a perfect pure For wave-alone flow, the near-bottom streaming is increasingly
current log-velocity profile with kn = 1.25 cm and u⁎c = 0.8 cm/s is offshore-directed as the bottom roughness increases, and this agrees
plotted in Fig. 20a. The component of the wave-induced mass transport qualitatively with Trowbridge & Madsen (1984). The mass transport
(here taken as the projection of the mass transport current obtained ex- flow also exhibits a near-linearly increasing velocity profile with eleva-
perimentally for orthogonal wave–current flows, i.e., similar to that tion for both smooth and rough bed experiments, and the agreement
shown in Fig. 13) in the direction of the external current is shown in with Longuet-Higgins (1953) solution deteriorates as the water surface
Fig. 20b. Combining these two current flows produces the velocity pro- is approached. Since experiments involving collinear wave–current
file depicted in Fig. 20a (denoted by hollow squares), showing a de- flows have shown that the near-surface mean flow is increased by op-
crease in the roughness inferred from the log-profile analysis relative posing waves (Klopman, 1994), incorporating this mechanism in a tur-
to the roughness obtained from the “non-contaminated” log profile. bulent wave-induced mass transport flow model appears needed in
Similar considerations of mean profile contamination by wave- order to improve our ability to predict this type of mean flow. The pres-
induced mass transport would lead to an overestimate of the bottom ent measurement may serve to validate such future models.
roughness for the 120° case, which also agrees with our findings With the addition of an orthogonal external turbulent current, the
(Table 3). This conjecture is supported by the results obtained by van wave-induced mass transport profile is transformed into a near-
Rijn & Havinga (1995), where the 60° case appears to have a larger uniform flow distribution. The resulting mass transport profile appears
near-bottom mean velocity and smaller apparent roughness compared to be more accurately described by Stokes' theory than the Longuet-
to the 120° case. Consequently, a new approach capable of isolating Higgins solution, except in the near-bed region where the boundary
the wave-induced mass transport from the external current profile, or layer effect prevails. Since the direction of mean flow is at 100° to the
a new boundary layer theory incorporating the effect of wave-induced x-axis, the mass transport profile observed in the near-uniform flow re-
mass transport on the wave–current mean flow, will need to be devel- gion is unlikely to originate from outside the channel, but instead is de-
oped before log-profile analysis of mean velocity profiles to resolve veloping from the inlet with a uniform flow distribution over depth.
the bottom roughness for near-orthogonal wave–current flows can be This is supported by the similar boundary layer thickness observed for
used with confidence. the mass transport flow profile and the external current flow that is
generated at the inlet, i.e. the mass transport current we observe has
6. Conclusions not reached full development over depth.
The superposition of waves on an orthogonal smooth turbulent cur-
A series of experiments on near-orthogonal combined wave–current rent increases the near-bottom mean velocity and reduces the bottom
flows have been performed at the Hydraulic Engineering Laboratory of roughness. The decrease in bottom roughness can be estimated with
NUS. Current-alone, wave-alone and combined wave–current flows the smooth turbulent roughness equation for steady flows, if the current
were generated over smooth and uniform marble-covered rough beds. shear velocity, u⁎c, is replaced by the maximum combined wave–current
Near-uniform flow regions were established in the current channels shear velocity, u⁎m, when the wave boundary layer is laminar and of a
for multiple measurements to be performed and to minimize uncertain- thickness, δw, smaller than the thickness of the viscous sublayer of the
ty associated with local bed irregularities. Velocity data were collected
over a duration of 3 min at each location to allow more than 100 wave
cycles to be measured during wave-alone and wave–current Table 3
kna, kn,cw and u⁎c for wave–current interaction over marble-covered bed
experiments.
The current-alone experiments show that the direction of mean Angle between waves and T kna Error u⁎c Error kn ,cw Error
flows are approximately the same as the desired flow directions created currents, ϕcw (deg) (s) (cm) (Factor) (cm/s) (±%) (cm) (Factor)
by the inlet and outlet alignments, i.e. at 60°, 90° and 120° to the direc-
1.4 3.61 1.42 0.91 10.83 0.48 1.74
tion (x) of the incident waves generated in the basin. The current expe- 60
1.6 3.51 1.80 0.99 11.44 0.47 2.23
riences a smooth turbulent roughness when it is generated over the 1.4 7.05 1.27 1.19 4.86 1.65 1.37
90
concrete bottom of the wave–current basin, while the bottom rough- 1.6 5.48 1.85 1.06 12.89 1.00 2.28
ness induced by a single layer of uniformly-arranged marbles is equiva- 1.4 7.06 1.56 1.10 5.54 1.68 1.67
120
1.6 8.09 1.40 1.08 17.17 1.43 1.78
lent to 2.4d50.
272 K.Y. Lim, O.S. Madsen / Coastal Engineering 116 (2016) 258–274

Fig. 20. (a) An example of a ‘perfect’ log-profile for an external current at 60° with the direction of the wave propagation (full circles) and the log-profile contaminated by mass transport
(hollow squares) (b) wave-induced mass transport profile projected onto the direction of the external current.

current, δvc, i.e. δw b δvc. For rough beds, the addition of waves to an or- Attempts to resolve the bottom roughness for the 60° and 120°
thogonal current leads to a reduction in the near-bottom mean velocity, wave-current cases have also been made. However, the contamination
as predicted by the Grant–Madsen (GM) model. However, both the ex- of the wave–current mean flow by the wave-induced mass transport
perimental results and the numerical simulation by Davies et al. (1988) current may lead to errors in estimating the bottom roughness using
indicate that the GM model has a tendency to over-predict the wave the log-profile analysis. Hence, it is not possible to obtain accurate
influence on near-orthogonal currents, in particular when the waves values of bottom roughness from the present 60° and 120° studies
are dominating the currents. A plausible explanation is that the wave until a new approach capable of isolating the wave-induced mass trans-
turbulence influencing the current flow is associated with the port from the external current, or a new boundary layer theory incorpo-
streamwise turbulence (u') when waves are collinear with the cur- rating the effect of wave-induced mass transport on the wave–current
rent, and spanwise turbulence (v') when waves are orthogonal to mean flow, is developed.
pffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffi
the current. Since v0 2 ̅ is considerably smaller than u0 2 ̅, the use
of the maximum wave–current shear velocity for all angles of Acknowledgment
wave–current interaction inevitably over-accounts for the wave ef-
fect on near-orthogonal currents, hence resulting in directional in- The authors gratefully acknowledge the financial support of the Na-
sensitivity of the GM model. A proposed approach to correct this tional Research Foundation of Singapore (NRF) through the Singapore-
weakness has been discussed and shown to improve the prediction MIT Alliance for Research and Technology's (SMART) Center for Envi-
of the bottom roughness, but further studies are required to ronmental Sensing and Modeling (CENSAM) program. The authors
strengthen the theoretical basis of this approach. In addition, other also thank Prof. Cheong Hin Fatt (NUS) for his helpful discussions and
wave–current models employing more advanced turbulence closure especially his assistance with lab infrastructure issues, and Prof. Alan
methods may also be validated and improved based on the results Davies and two other reviewers for their insightful comments and sug-
and insights obtained in the present study. gestions on the paper.
As shown by Davies et al. (1988) and Afzal et al. (2015), the mean
current flow for combined non-collinear waves and current should Appendix A
veer towards the direction normal to the direction of the near-bottom
wave orbital velocity as the bottom is approached from above. However, A two-layer time-invariant, linearly-varying eddy viscosity distribu-
this wave direction, expected to coincide with the x-axis, was locally af- tion was adopted by Grant & Madsen (GM) in their wave–current
fected by the presence of parasitic waves emanating from the current boundary layer model (Grant & Madsen, 1979, 1986). Within the
inlet and outlet. Within the region of measurement these parasitic wave boundary layer, the eddy viscosity is scaled by the maximum com-
waves were shown to form a partially standing wave in the direction bined wave–current shear velocity, u⁎m, whereas outside of it, the eddy
of the nominal current channel and therefore introduce unexpected viscosity is solely characterized by the current shear velocity, u⁎c:
wave velocities in the y-direction. Resolving the parasitic wave charac- υcw ¼ κ ðum Þz f or z ≤δcw ðA:1Þ
teristics in our experiments, the resulting effective wave direction was
found to potentially vary by ± 10° around the x-axis. In the 60° and υc ¼ κ ðuc Þz f or zNδcw ðA:2Þ
120° flow experiments, the nominal angle between current and the nor-
mal to the incident waves is ±30° and the parasitic wave effect merely where δcw = combined wave–current boundary layer thickness, i.e. the
changes the magnitude of the veering angle but not its direction, and hypothetical elevation at which the two logarithmic velocity profiles
our observed veering is in qualitative agreement with Davies et al.'s (one within and one outside the wave boundary layer) intersect. Com-
(1988) prediction. However, for the 90° case, the direction of near- putational procedures of the GM model are shown as follows:
bottom mean velocity veering was found to be significantly affected
by the presence of parasitic waves and to depend on the local effective (1) Calculate the ratio of the current shear stress to the maximum
wave direction. This finding highlights the potential importance of rec- wave shear stress, μ:
ognizing the existence of parasitic waves in experimental set-up
employing gravity fed current generation and consider their presence τc
μ¼ ðA:3Þ
in experimental design and data analysis. τwm
K.Y. Lim, O.S. Madsen / Coastal Engineering 116 (2016) 258–274 273

where τc is the, assumed known, current shear stress and τwm is the (6) Predict apparent roughness, kna.
maximum wave shear stress. Since τwm is unknown, the initial value
of μ is taken as zero (assuming τwm NN τc).
The velocity profiles in the GM model are expressed as follows:
(2) Obtain the enhancement factor, Cμ:
 
uc uc z
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi u¼ ln f or z ≤δcw ðA:12Þ
um κ z0
τ m ¼ τwm 2 þ 2τc τwm j cosϕcw j þ τ c 2 ðA:4Þ
¼ C μ τwm  
uc z
u¼ ln f or zNδcw ðA:13Þ
κ z0a
where τm is the vector summation of the wave and current shear stress
aligned at an angle of ϕcw, and: where kna = 30zoa is the apparent roughness, and equating the above
two formulas at z = δcw yields:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Cμ ¼ 1 þ 2μ j cosϕcw j þ μ 2 ðA:5Þ  uc =um
z0
zoa ¼ δcw : ðA:14Þ
δcw
(3) Compute the combined wave–current friction factor, fcw
(Madsen, 1994):
References
"  #
C μ Abm −0:078 Afzal, M.S., Holmedal, L.E., Myrhaug, D., 2015. Three-dimensional streaming in the seabed
f cw ¼ C μ exp 7:02 −8:82 ðA:6Þ boundary layer beneath propagating waves with an angle of attack on the current.
kn J. Geophys. Res. Ocean, 120, 4370–4391.
Arnskov, M.M., Fredsøe, J., Sumer, B.M., 1993. Bed shear stress measurements over a
smooth bed in three-dimensional wave–current motion. Coast. Eng. 20, 277–316.
C μ Abm Barrantes, A.I., Madsen, O.S., 2000. Near-bottom flow and flow resistance for currents
for 0:02b kn
b102 obliquely incident to two-dimensional roughness elements. J. Geophys. Res. 105
(C11), 26,253–26,264.
"  # Carstens, M.R., Neilson, F.M., Altinbilek, H.D., 1969. Bed forms generated in the laboratory
C μ Abm −0:109 under an oscillatory flow: analytical and experimental study. US Army Corps of Engi-
f cw ¼ C μ exp 5:61 −7:30 ðA:7Þ neers, Technical Memorandum no. 28.
kn Chanson, H., 2008. Acoustic doppler velocimetry (ADV) in the field and in laboratory:
practical experiences. International Meeting on Measurements and Hydraulics of
Sewers IMMHS '08, Bouguenais, France, pp. 49–66.
C μ Abm Christofferson, J.B., Jonsson, I.G., 1985. Bed friction and dissipation in a combined current
for 102 b kn
b104 and wave motion. Ocean Eng. 12 (5), 387–423.
Coffey, F.C., Nielsen, P., 1986. The influence of waves on current profiles. Proc. 20th Inter-
national Conference on Coastal Engineering, pp. 82–96.
where Abm is the near-bottom particle excursion amplitude and kn = kn,cw
Davies, A.G., Villaret, C., 1997. Oscillatory flow over rippled beds: Boundary layer structure
is the equivalent Nikuradse sand grain roughness of the bottom. and wave-induced Eulerian drift. Chapter 6 in gravity waves in water of finite depth.
In: Hunt, J.N. (Ed.), Advances in Fluid Mechanics. Computational Mechanics Publica-
(4) Obtain the maximum wave shear stress, τwm: tions, pp. 215–254.
Davies, A.G., Villaret, C., 1999. Eulerian drift induced by progressive waves above rippled
and very rough beds. J. Geophys. Res. 104 (C1), 1465–1488.
1 Davies, A.G., Soulsby, R.L., King, H.L., 1988. A numerical model of the combined wave and
τwm ¼ f ρu 2 ðA:8Þ
2 cw bm current bottom boundary layer. J. Geophys. Res. 93 (C1), 491–508.
Fernando, P.C., Guo, J., Lin, P., 2011. Wave–current interaction at an angle 1: experiment.
J. Hydraul. Res. 49 (4), 424–436.
where ubm is the near-bottom wave orbital velocity amplitude and ρ is Fredsøe, J., 1984. Turbulent boundary layer in wave–current motion. J. Hydraul. Eng. ASCE
110, 1103–1120.
the fluid density. τwm is substituted back into (A.3) to obtain a new value
Fredsøe, J., Andersen, K.H., Sumer, B.M., 1999. Wave plus current over a ripple covered
of μ, and Eqs. (A.3) to (A.8) are iterated until the convergence of μ. bed. Coast. Eng. 38, 177–221.
Grant, W.D., Madsen, O.S., 1979. Combined wave and current interaction with a rough
(5) Compute the transition level for the two-layer eddy viscosity, bottom. J. Geophys. Res. 84 (C4), 1797–1808.
δcw: Grant, W.D., Madsen, O.S., 1986. The continental shelf bottom boundary layer. Annu. Rev.
Fluid Mech. 18, 265–305.
 P Grass, A.J., 1971. Structural features of turbulent flow over smooth and rough boundaries.
κum Cμ 1 J. Fluid Mech. 50 (2), 233–255.
δcw ¼ A ðA:9Þ
ω μ 16:3 George, W.K., 2007. Is there a universal log law for turbulent wall-bounded flows? Phil.
Trans. R. Soc. A 365, 789–806.
Havinga, F.J., 1992. Sediment concentration and sediment transport in case of irregular
where, non- breaking waves with a current. Part E. Technical Report. Delft University of
Technology, Netherlands.
"   # Holmedal, L.E., Johari, J., Myrhaug, D., 2013. The seabed boundary layer beneath waves
C μ Abm −0:071 opposing and following a current. Cont. Shelf Res. 65, 27–44.
A ¼ exp 2:96 −1:45 ðA:10Þ
kn Huang, Z., Mei, C.C., 2003. Effects of surface waves on a turbulent current over a smooth or
rough seabed. J. Fluid Mech. 497, 253–287.
pffiffiffiffiffiffi Jackson, P.S., 1981. On the displacement height in the logarithmic velocity profile. J. Fluid
Cμ Mech. 111, 15–25.
P¼ pffiffiffiffiffiffi pffiffiffi ðA:11Þ Johnson, J.W., 1953. Engineering aspects of diffraction and refraction. Trans. Am. Soc. Civ.
2 Cμ − μ
Eng. 118 (2556), 617–652.
Jonsson, I.G. 1966. Wave boundary layers and friction factors. Proc. 10th International
It should be noted that Eqs. (A.9) through (A.11) differ from the orig- Coastal Engineering Conference, ASCE, 1, 127–148.
Kemp, P.H., Simons, R.R., 1982. The interaction between waves and a turbulent current:
inal GM model formulation proposed by Grant & Madsen (1979, 1986).
waves propagating with the current. J. Fluid Mech. 116, 227–250.
The derivation of Eq (A.10) was given by Madsen & Salles (1998), while Kemp, P.H., Simons, R.R., 1983. The interaction between waves and a turbulent current:
the additional factor of (Cμ/μ)P/16.3 in Eq. (A.11), which is typically close waves propagating against the current. J. Fluid Mech. 130, 73–89.
to 1/3, is added to obtain an improved transitional level between the Klebanoff, P.S., 1955. Characteristics of turbulence in a boundary layer with zero pressure
gradient. NACA Rep. 1247.
two logarithmic velocity profiles derived from the eddy viscosity formu- Klopman, G., 1994. Vertical structure of the flow due to waves and currents. Delft Hydrau-
lation given in Eqs. (A.1) and (A.2), respectively. lics Progress Report H 840.30, Part II.
274 K.Y. Lim, O.S. Madsen / Coastal Engineering 116 (2016) 258–274

Lim, K.Y., 2013. An experimental study on mean flow characteristics for near-orthogonal Myrhaug, D., Slattelid, O.H., 1989. Combined wave and current boundary layer model for
combined waves and currents. Ph.D. Thesis. Department of Civil & Environmental En- fixed rough seabeds. Ocean Eng. 16 (2), 119–142.
gineering, National University of Singapore. Nadaoka, K., Kondoh, T., 1982. Laboratory measurements of velocity field structure in the
Lim, K.Y., Madsen, O.S., Cheong, H.F., 2012. Current characteristics in the presence of near- surf-zone by LDV. Coast. Eng. Jpn 25, 125–145.
orthogonal waves. Proc. 33rd International Conference on Coastal Engineering, San- Nielsen, P., 1992. Coastal Bottom Boundary Layers and Sediment Transport. World
tander, Spain, Paper No.: Currents.41. Scientific, Singapore.
Lofquist, K.E.B., 1986. Drag on naturally rippled beds under oscillatory flows. Technical Olabarrieta, M., Medina, R., Castanedo, S., 2010. Effect of wave–current interaction on the
Report No. MP-86-13, U.S. Army Corps of Engineers. Coastal Engineering Research current profile. Coast. Eng. 57, 643–655.
Center. Pham-Gia, T., Hung, T.L., 2001. The mean and median absolute deviations. Mathematical
Longuet-Higgins, M.S., 1953. Mass transport in water waves. Philos. Trans. R. Soc. London, and Computer Modeling 34 (7–8), 921–936.
Ser. A 245, 535–581. Scandura, P., Foti, E., 2011. Measurements of wave-induced steady currents outside the
Lundgren, H., 1972. Turbulent currents in the presence of waves. Proc. 13th International surf zone. J. Hydraul. Res. 49 (sup1), 64–71.
Coastal Engineering Conference. ASCE, pp. 623–634. Schlichting, H., 1968. Boundary Layer Theory. sixth ed. McGraw-Hill, New York.
Madsen, O.S., 1970. Waves generated by a piston-type wave maker. Proc. 12th Interna- Sleath, J.F.A., 1987. Turbulent oscillatory flow over rough beds. J. Fluid Mech. 182,
tional Conference on Coastal Engineering, ASCE 1, Washington, DC, pp. 589–607. 369–409.
Madsen, O.S., 1994. Spectral wave–current bottom boundary layer flows. Proc. 24th Inter- Simons, R.R., Grass, T.J., Mansour-Tehrani, M., 1992. Bottom shear stresses in the bound-
national Conference on Coastal Engineering. ASCE, pp. 384–398. ary layer under waves and currents crossing at right angles. Proc. 23rd Int. Conf. on
Madsen, O.S., 2009. Lecture notes. Ralph M. Parsons Laboratory, Massachusetts Institute of Coastal Eng. 45, 1, pp. 604–617.
Technology, Cambridge, M.A. Smith, J.D., 1977. Modeling of sediment transport on continental shelves. The Sea 6. Wiley
Madsen, O.S., Kularatne, S., Cheong, H.F., 2008. Experiments on bottom roughness experi- Interscience, New York, pp. 539–577.
enced by currents perpendicular to waves. Proc. 31st International Conference on Soulsby, R.L., Hamm, L., Klopman, G., Myrhaug, D., Simons, R.R., Thomas, G.P., 1993.
Coastal Engineering, ASCE. 1, pp. 845–853. Wave–current interaction within and outside the bottom boundary layer. Coast.
Madsen, O.S., Salles, P., 1998. Eddy viscosity models for wave boundary layers. Proc. 26th Eng. 21, 41–69.
International Conference on Coastal Engineering, ASCE 3, pp. 2615–2627. Trowbridge, J.H., Madsen, O.S., 1984. Turbulent wave boundary layers: 2. Second-order
Mathisen, P.P., Madsen, O.S., 1996a. Waves and currents over a fixed rippled bed: 1. Bot- theory and mass transport. J. Geophys. Res. 89 (C5), 7999–8007.
tom roughness experienced by waves in the presence and absence of currents. Van Rijn, L.C., Havinga, F.J., 1995. Transport of fine sands by currents and waves II. J. Wa-
J. Geophys. Res. 101 (C7), 16533–16542. terway, Port, Coastal and Ocean Eng. 121 (2). ASCE, pp. 123–133
Mathisen, P.P., Madsen, O.S., 1996b. Waves and currents over a fixed rippled bed: 2. Bot- Yuan, J., Madsen, O.S., Chan, E.S., 2012. Experimental study of turbulent oscillatory bound-
tom and apparent roughness experienced by currents in the presence of waves. ary layers in a new oscillatory water tunnel. Proc. 33rd International Conference on
J. Geophys. Res. 101 (C7), 16543–16550. Coastal Engineering, ASCE. Paper No.: Waves.24.
Musumeci, R.E., Cavallaro, L., Foti, E., Scandura, P., Blondeaux, P., 2006. Waves plus cur- Yuan, J., Madsen, O.S., 2014. Experimental study of turbulent oscillatory boundary layers
rents crossing at a right angle: experimental investigation. J. Geophys. Res. 111, in an oscillating water tunnel. Coast. Eng. 89, 63–84.
C07019. http://dx.doi.org/10.1029/2005JC002933.

You might also like