2406.00641v1
2406.00641v1
2406.00641v1
4 Institute of Port, Coastal and Offshore Engineering, Zhejiang University, Zhejiang, China.
7 The Key Laboratory of Water and Sediment Sciences, Ministry of Education, Peking University,
Beijing, China
Highlights
-Coherent airflow sweep structures exerting sufficient drag force can set coarse particles in motion.
-A work-based criterion has been established to define the full aerodynamic entrainment of coarse
particles.
1
Abstract
The role of coherent airflow structures capable of setting gravel-size particles in motion is studied
is proposed to describe the incipient motion of large-particles ranging from rocking (incomplete
entrainment) to incipient rolling (full entrainment). Wind tunnel experiments were conducted on an
aerodynamically rough bed surface under near-threshold airflow conditions. Synchronous signals of
airflow velocities upwind of the test particles and particle displacement are measured using a hot film
anemometer and a laser distance sensor, respectively, from which coherent airflow structures
(extracted via quadrant analysis) and particle movements are interlinked. It is suggested that the
incipient motion of gravel-size particles (rocking and rolling) may result from sufficiently energetic
sweep events corresponding to aerodynamic drag forces in excess of the local micro-topography
resistance. However, full entrainment in rolling mode should satisfy the presented work-based criterion.
Furthermore, using an appropriate probabilistic frame, the proposed criterion may be suitable for
describing processes of energy transfer from the wind to the granular soil surface, ranging from the
creep transport of gravels to the “mechanical sieving” of mega-ripples, as well as the transport of light
The entrainment of particles from a particle bed is a key problem in wind-driven sediment transport.
Most existing models define critical entrainment conditions as those at which a bed particle begins to
move. Such conditions are typically met during the passing of turbulent flow structures. However, the
energy transferred from the structure to the particle may not be sufficient to leapfrog over neighboring
2
bed grains. In this case, the particle will eventually fall back into its bed pocket. Here, we
experimentally evaluate an alternative work-based criterion for particle entrainment during the passing
of turbulent flow structures. Our results indicate that sweeps of sufficient turbulent energy are
Keywords: quadrant event; particle rocking; incipient rolling; incipient motion; threshold energetic
1. Introduction
The creep of individual gravel particles is sporadic and intermittent, as demonstrated by tracking the
motion of gravel on the crest of gravel mega-ripples in Wright Valley, Antarctica (Gillies et al., 2012)
and the Argentinean Puna (de Silva et al., 2013). Such movement is a key process in the formation and
evolution of coarse-grained bed forms (Zimbelman et al., 2009; Milana, 2009; Yizhaq et al., 2012).
Coarse gravel particles form a surface layer, effectively protecting mega-ripples from wind erosion
(Sharp, 1963). However, the sizes of such mega-ripples are limited due to their migration, which may
be initiated either by strong wind gusts or by impinging particles transported in saltation mode
(Isenberg et al., 2011, Katra et al., 2014, Tholen et al., 2022). In addition to the aforementioned low
mobility transport mode of gravel, other potential mechanisms of transfer of wind energy towards
performing geomorphic work have been recently identified (de Silva et al., 2013). Particularly in the
case where the local micro-topography prevents particles from moving downwind, the resulting
rocking motion induced by turbulent airflow might be instrumental for the growth of gravel mega-
ripples through "mechanical sieving," a process suggested by de Silva et al. (2013). Thus, a better
3
understanding of the formation and evolution of coarse-grained bed forms, for instance, for modeling
purposes, requires detailed investigations of intermittent creep and rocking of gravel particles
Traditionally, the unsteady wind is treated as a fully random motion in modeling sediment
transport (Anderson, 1987; Spies and McEwan, 2000; Lu et al., 2005; Li et al., 2008, Kok and Renno,
2009). However, studies based on flow visualization have noted that wall-bound turbulence consists
of random fluctuations and intermittent and quasi-periodic coherent flow structures (e.g., Kline et al.,
1967; Grass, 1971; Adrian, 2007). Such turbulent flow structures (Robinson, 1991) or coherent flow
structures (Bauer et al., 2013) hold considerable promise for providing new insights into the mechanics
of sediment transport and evolution of bedforms because they generate locally high Reynolds stress,
which may account for the heterogeneity of sediment transport across a hierarchy of temporal and
There is now an increasing understanding that coherent airflow structures may be a major
contributor to the aeolian transport of relatively small particles, e.g., sand grains (Bauer et al., 1998;
Schönfeldt and von Löwis, 2003; Baas and Sherman, 2005; Chapman et al., 2012, Pähtz et al., 2018).
Specifically, flow structures associated with the bursting-like phenomenon (sweeps and ejections) have
been of considerable significance for transport by saltation (Sterk et al., 1998; Leenders et al., 2005;
Wiggs and Weaver, 2012). However, very few studies report the role of flow structures in the
intermittent creep of large-size sediments. This episodic transport of bed material may be inherited
from the intermittent character of coherent structures of wind gusts found in natural aeolian
environments (e.g., van Boxel et al., 2004; Walker, 2005), which locally disrupt the laminar boundary
4
Generally, because the carrying capacity for water is greater than that for air, the transport of
coarse bed material, typically larger than fine gravel, has received more attention in water
environments (e.g., Drake et al., 1988; Thorne et al., 1989; Hardy et al., 2009). Investigations show
that the incipient rolling of particles is associated with intense sweep events characterized by an
optimum combination of fluid force and duration (Valyrakis et al., 2011a; Dwivedi et al., 2011). These
sweep events may be associated with the motion of large-scale retrograde vortices (Wu and Shih, 2012)
or octant structures (Keylock et al., 2014). Even though significant progress has been made in
identifying the dynamic processes leading to the entrainment of large sediment particles (Pähtz et al.,
2020), this has not been met in the case of aeolian transport.
This study aims to thoroughly investigate the role of coherent airflow structures on the incipient
motion of an individual large (typically larger than fine gravel) particle exposed on a bed surface
between the flow structures responsible for rocking (incomplete entrainment) and incipient rolling (full
entrainment). This is achieved by considering an energy transfer framework from airflow structures to
surface particles for the particle-scale micro-mechanical model. In addition, wind tunnel experiments
involving synchronous measurements of local airflow velocity and particle incipient displacements are
conducted near the threshold flow condition to verify the proposed model. Considering the potential
significance of coherent flow structures on aeolian transport and the lack of comprehensive studies for
the aeolian entrainment of gravel particles, this study may be seen as a novel paradigm to
comprehensively describe the mechanisms of intermittent geomorphic work of the soil surface due to
5
2. Modeling the role of coherent flow structures on incipient motion
Near-threshold (with regard to particle transport) flow conditions are considered, during which particle
motion occurs intermittently. In the case of the simplified micro-mechanical model shown in Fig. 1, it
is possible that relatively short-lived but sufficiently energetic coherent flow structures, which
energetic retrograde vortices may represent following Wu and Shih (2012), can overcome the intrinsic
resistance for this arrangement, setting the exposed particle into motion. At near-threshold conditions,
entrainment is a dynamic and sporadically occurring process that cannot be modeled suitably by the
average wind strength, with spatially and temporally averaged measures such as frictional velocity
(e.g., Greeley and Iversen, 1985; Shao and Lu, 2000). An approximate mechanical model should
consider interactions between the flow structures and the individual solid particle.
Fig. 1. Sketch of a simplified micromechanical model of an exposed particle resting on a rough surface
comprising of same-sized particles, showing the energy transferred from the wind burst to the particle,
6
CeffEf, and the specific elevation (zcr) beyond which entrainment is possible by the action of the mean
flow forcing alone. Ef is the energy content of the flow structure, and Ceff is the energy transfer
coefficient.
To investigate the above postulation, the following framework may be considered based on the
micro-mechanical arrangement shown in Fig. 1. It is hypothesized that the particle movement from the
resting position to a specific elevated position beyond which entrainment is possible by the action of
the mean flow forcing alone (O to Oʹ, as shown in Fig. 1) may be initiated by a sufficiently energetic
passing flow structure. Based on the conservation of energy, the energy the particle obtains CeffEf
originates from the energy content of the flow structure Ef with a transfer coefficient Ceff similar to
Valyrakis et al. (2013). This process is achieved by mechanical work of aerodynamic force induced by
the flow structure. Then, an appropriate criterion for specifying the threshold level for particle
entrainment should link the amount of energy offered by the flow structure with the minimum amount
To overcome the energy barrier from O to Oʹ, the mechanical work exerted on the particle should be
at least in excess of the increase of gravitational potential (Fgzcr; zcr is the vertical displacement from
O to O’):
where Fg [=πd3(ρp−ρa)g/6] is the effective gravity of the particle, d is the particle diameter, ρp and ρa
are, respectively, the density of the particle and air, and g is the gravitational acceleration constant.
Rolling friction forces can be two orders of magnitude smaller than the particle’s weight (Pöschel et
al. 1999). For Eq.(2), any energy losses due to rolling frictional forces are significantly smaller than
7
the uncertainties in calculating the particle and flow event energies and thus can safely be neglected.
Substituting Eq. (2) into Eq. (1) and linking zcr to the particle diameter (d) using a parameter α (zcr=αd)
yields:
The energy density of individual coherent flow structures may be sufficiently represented with the
cube of the local and instantaneous windwise velocity component upwind of particles (u3), neglecting
the contribution of any other components (similar to Valyrakis et al. 2013 for water flows). Then Ef
can be expressed as the integral of energy density over the time scale (or duration) of the flow structure
where t0 is the occurrence time of flow structures, and S (=0.25πd2) is assumed to be the spherical
particle's cross-sectional area. Substituting Eq. (4) into Eq. (3) yields:
′ 𝑡 +𝑇𝑓
𝐶𝑒𝑓𝑓 ∫𝑡
0
𝑢3 𝑑𝑡 > 𝜔2 𝑑𝛼 (5)
0
where C'eff (=Ceff/Cd) is the normalized energy transfer coefficient, Cd is the drag coefficient, and 𝜔 =
√4𝑔𝑑(𝜌𝑝 − 𝜌𝑎 )/(3𝐶𝑑 𝜌𝑎 )) is the fall velocity of particles in air (Durán et al., 2011). Eq. (5) defines a
work-based criterion for full entrainment of coarse particles by rolling mode. It links the threshold
energy level relevant to flow structures to the properties of solid particles and flow reflected by particle
fall velocity ω and a specific elevation parameter α which may be a random variable at least associated
′ 𝑡 +𝑇𝑓
The amount of energy offered from flow events to particles, parameterized 𝐶𝑒𝑓𝑓 ∫𝑡
0
𝑢3 𝑑𝑡
0
according to Eq. (5), is equal to the mechanical work performed on particles neglecting the frictional
energy loss. Without considering lift force on particles, the mechanical work is integral to the rate of
8
work of the drag force. The drag force is represented using the classical quadratic velocity
parameterization (u2), neglecting the particle’s entrainment velocity. Then, similar to Valyrakis et al.
(2013), the rate of work (product of drag force and particle windwise velocity v) can be parameterized
by u2v, and the mechanical work done from the flow structures on the rolling particle can be
𝑡 +𝑇𝑓 ′ 𝑡 +𝑇𝑓 𝑡 +𝑇𝑓
parameterized using ∫𝑡 0 𝑢2 𝑣𝑑𝑡 the above analyses to develop 𝐶𝑒𝑓𝑓 ∫𝑡
0
𝑢3 𝑑𝑡 = ∫𝑡 0 𝑢2 𝑣𝑑𝑡
0 0 0
the equation for estimating the normalized energy transfer coefficient, C′eff :
′ 𝑡 +𝑇𝑓 𝑡 +𝑇𝑓
𝐶𝑒𝑓𝑓 = ∫𝑡 0 𝑢2 𝑣𝑑𝑡 / ∫𝑡 0 𝑢3 𝑑𝑡 (6)
0 0
Characteristic values or stochastic distributions of particle sizes and bed configurations can be
determined for a specific granular surface. In this case, for a flat, fixed, and well-packed bed surface
comprising of unisize spherical particles, parameter α from Eq. (5) is expected to remain invariant and
can be defined experimentally. Further, the normalized energy transfer coefficient C′eff defined by Eq.
3. Experiments
The experiments were undertaken in an environmental wind tunnel in the State Key Joint Laboratory
of Environmental Simulation and Pollution Control, Peking University. The available rectangular
section of the wind tunnel is 30 m long, 3 m wide, and 2 m high. The oncoming wind speed (U)
measured in the middle of the tunnel cross-section, 0.4 m away from the tunnel entrance and 1.2 m
above the tunnel floor, can be varied continuously from 0.5 to 20 m/s.
The complexities associated with the variability of natural grains were eliminated using hollow,
spherical particles of d=40 mm weighing 2.7 g (e.g., a typical table tennis ball made of celluloid) for
9
the test particle. The wind tunnel's bed surface at the test section comprises one layer (bed depth of 4
cm) of uniformly arranged (in a rectilinear fashion), unisize particles of the same properties as the test
particle, towards forming a flat and aerodynamically roughened experimental bed surface. The test
surface is non-erodible for the tested wind speeds. The fixed bed surface allows for standard
estimations of the resistance to entrainment with average shear stress criteria, which, however, have
been demonstrated to overestimate the mobility of the bed surface material (Yager et al., 2018; Pähtz
et al., 2020). The 4 m long, 2.4 m wide experimental bed test section was located 22.6 m downwind
from the tunnel’s entrance, where fully developed turbulent conditions were established during the
experiments.
Several factors were considered in selecting the particle size. Specifically, the particle size should
be a few cm at least to achieve the smallest relative error when considering the resolution of the optical
method (laser distance meter with ±0.1 mm accuracy) used in the experiments. Likewise, as a rule of
thumb, the chosen particle size should be a fraction (e.g., less than 1/20th) of the effective width of the
wind tunnel (which is the portion of the width of the 3 m wide tunnel along which the airflow is fully
Different particle property combinations were tested. For example, greater size spherical particles
might also be able to be entrained even if they might be heavier because they impinge higher into the
boundary layer. However, a size and weight combination that would allow the particle to be entrained
by the fluctuating coherent flow structures impinging into or generated near the bed surface, rather
than because of the average aerodynamic forcing experienced further from the bed surface, is strongly
preferred. This would allow the reproduction of the highly intermittent character of aerodynamic
entrainment, also seen in the field. Last, the combination of the target particle’s weight and size should
10
be such that it would allow for aerodynamic entrainment to be observed for a range of wind tunnel
speeds that the turbine can generate. This consideration automatically excludes the ranges of heavier
If natural gravel with a particle size similar to the test particles were employed in these
experiments, a high-speed wind tunnel with wind strength of up to approximately 100 m/s would be
required (Batt and Peabody, 1999). Therefore, low-density particles (about 67 times air density) were
utilized to allow the running of the experiments in the available low-speed wind tunnel. A similar
approach has been used in hydraulic transport for a long period, e.g., Kaftori et al. (1995) used
polystyrene particles to investigate the role of coherent wall structures on sediment transport in a
laboratory flume. Even though the experimental particles are lighter than natural sediments, they are
much heavier than air. Then, the incipient motion of the model is still governed by a self-weight and
aerodynamic force similar to natural gravel. Thus, this experimental setup may be appropriate to
investigate the elementary dynamics of entrainment for ideally spherical gravel particles on uniform
bed surfaces.
the environment due to wind. The current analysis focuses on wind tunnel experiments that
demonstrate for a first time the assessment of the aerodynamic threshold which accounts for turbulence
for the incipient motion of plastic debris, albeit with an idealised spherical shape. These are considering
the new dynamic criterion of energy, which fully accounts for the effects of turbulent fluctuations, that
11
Fig. 2. Illustration of the experimental setup used for conducting synchronous particle displacement
and airflow velocity measurements. The top left insert, shows the target particle in its resting
tetrahedral arrangement, along with an indication of the range of the measurements taken by the laser
distance sensor. The bottom left insert, shows the target particle rocking outside of its resting pocket,
Movements of the particles were recorded by a laser distance sensor (LDS) with 0.1 mm resolution.
Calibration revealed that the electrical signal output from the LDS was linear with the distance from
the target object to the receptor (similar to Diplas et al., 2010 and Xiao-Hu et al., 2021). The LDS laser
beam was fixed to shine at the center of the downstream face of the target particle (Fig. 2). The sensor's
measuring range of interest was set from the rest position of the target particle to the top of the
supporting downwind particles, where the mobile particle would be considered to have been
completely entrained. The LDS measurement range of interest refers to the maximum distance LDS
12
records when the target particle is entrained from the resting position to the position, signifying that
full entrainment has been achieved. Continuous measurement records are allowed by using a pin that
forces the particle back into its pocket after the fluctuating aerodynamic forcing from advected airflow
structures is reduced below the particle’s resisting force (which is a component of the particle’s weight).
For the instances when the aerodynamic forces are too strong, the particle may dislodge further
downwind past the pin (Fig. 2). Since only the particle’s incipient motion is studied, only records
within the LDS measurement range of interest are kept for further analysis. Utilizing such a setup and
considering the fixed pocket geometry allows the measured displacement to be decomposed into its
The instantaneous windwise and vertical components of wind velocities (u and w, respectively)
were measured using a two-dimensional hot-film anemometer positioned at one diameter (1d) upwind
of the front of the particle and 0.2d above the top of the target particle in its resting pocket position
(Fig. 2). The velocity measurement location was chosen in order to avoid being too close to the particle,
potentially interfering with the local flow field around the particle, or too far, where the measurements
would not be representative of the local aerodynamic forcing field around the test particle. Signals of
the LDS and the anemometer were sampled synchronously using a multichannel data acquisition card
(DAQ) with a frequency of 1000 Hz. Further information on the acquisition and processing of the
The approach velocity (U) upwind of the test section in the experiment remained relatively invariant
at 8 m/s near the threshold airflow condition. The average windwise wind velocity upwind of the target
particle (𝑢̅) was approximately 4.8 m/s, and the corresponding particle Reynolds number (𝑅𝑒𝑝 = 𝑢̅𝑑/𝑣,
13
where ν is the kinematic viscosity of air and ν=1.5×10-5 m2/s) was approximately 13,000. At such a
high Reynolds number, the airflow around particles is fully turbulent, and the drag force results from
the asymmetry of pressure between the two sides of the particle. This regime is consistent with that of
natural gravels, considering their large size, density, and the resulting high fluid threshold.
The conditions of the simulated atmospheric boundary layer are shown in Fig. 3, where the profile
of the average windwise wind velocity at the measuring location followed the typical logarithmic law
10 mm above the tops of the particles comprising the bed surface (Fig. 3a). The turbulence intensity is
decreased with the height from the bed surface, with a value of up to 20% near the bed surface (Fig.
3b). The inertial subrange at frequencies >10 Hz for windwise velocity components upwind of the
particle is characterized by a typical Kolmogorov −5/3 power law (Fig. 3c). The energy-containing
range at frequencies <10 Hz, scaled with a −1 spectral slope, is ascribed to large-scale eddies in which
the energy production and cascade energy transfer coexist. These results are in accordance with a
Fig. 3. (a) Windwise wind velocity profile, (b) turbulence intensity profile, and (c) u-power spectra at
the measuring location at U=8 m/s. The zero-reference level is set to the tops of the rough bed's
14
3.3. Data processing
Coherent flow structures associated with near-surface bursting processes were extracted using the
quadrant technique given by Lu and Willmarth (1973). Briefly, the coherency of local airflow is
identified by the magnitude of kinematic Reynolds stress (negative covariance of u' and w', –u'w')
based on a subjective threshold |𝑢′ 𝑤 ′ | ≥ 𝐻𝑢𝑟𝑚𝑠 𝑤𝑟𝑚𝑠 , where urms and wrms are the root-mean-square
values of u' and w' signals, respectively. The parameter H in this study was set to 1, following the
detailed investigation of Bogard and Tiederman (1986). Stress events of high magnitude, which
indicate the presence of coherency in the local flow, are decomposed into four categories according to
u' and w' quadrants: i) quadrant 1 (Q1) outward interactions (u'>0, w'>0), ii) quadrant 2 (Q2) ejections
(u'<0, w'>0), iii) quadrant 3 (Q3) inward interactions (u'<0, w'<0), and iv) quadrant 4 (Q4) sweeps
(u'>0, w'<0). Stress events below the threshold are called “hole events” comprising noncoherent (or
Following the above procedure, synchronous time series of quadrant and displacement events are
acquired and subsequently analyzed. Signatures of quadrant events are dominated by Q4 and Q2 events,
which occur with a frequency of approximately 2 counts/s as shown in Table 1. Q1 and Q3 events
infrequently occur with a frequency of 0.2 and 0.3 counts/s, respectively. Correspondingly, the average
value of interval time (Ti) of Q1 and Q3 events is much higher than that of Q2 and Q4 events. The
location of the anemometer probe may influence the detection of the occurrence of quadrant events.
According to Wu and Shih (2012), if the probe is positioned relatively far (e.g., one particle diameter
similar to our experimental setup) from the upwind front of the target particle, the quadrant technique
is inclined to predict more Q4 events than Q1 events. This might be true considering the body of flow
structures is three-dimensional, and a single-point velocity measurement can only grasp the local
15
features of flow structures. The displacement events occur much less frequently than quadrant events
(see Table 1). This suggests that most quadrant events are ineffective for particle incipient motions if
the latter is supposed to result from quadrant events. The average value of the duration of quadrant
events (Tf) is smaller than the average value of ascending duration of particles (ta), defined as the
duration of downwind movements from the resting location to the peak for rocking motions and to the
location where full entrainment is achieved for incipient rolling. This may imply that quadrant events
Table 1 Time-averaged quadrant and displacement event parameters, including frequency, duration,
and interval time (until the stochastically occurring full entrainment by rolling, when the test particle
4. Results
The test particle was observed to roll intermittently and to rock preceding the incipient rolling under
our experimental conditions, similar to the behavior of natural gravels (de Silva et al., 2013). To
examine which quadrant event was responsible for each recorded particle motion, comparisons were
made of the occurrence of rocking and rolling to quadrant events. It is found that the rocking motions
can be caused by Q4 or Q1 events alone (e.g., event A in Fig. 4a and event B in Fig. 4b, respectively)
or by sequences of Q1 events followed by Q4 events (e g., event C and event D in Fig. 5a). The
16
occurrence of incipient rolling similarly has an intimate relationship with sweep events as event E
shown in Fig. 5b. Descriptive statistical indices for all movements detected in the experiments, such
as counts of rocking (incomplete particle motions) and rolling (full entrainments) and the type of flow
structure responsive for this particle response, are reported in Table 2. It is worth noting that the
majority of complete and partial particle displacements, more than 85%, occur due to Q4 events alone.
Thus, sweeps are the most relevant quadrant events for the incipient motion of coarse particles. These
results agree with some investigations in water flows (e.g., Hofland and Booij, 2004; Detert et al.,
2010). However, they are distinct from what Wu and Shi (2012) reported, suggesting that Q1 events
dominate the instant of particle entrainment. This may be because the measuring location of flow
velocity in Wu and Shi (2012) was very close to the target particle (1/8d), and Q4 events were not
17
Fig. 4. Representation of 1s histories of quadrant events (described respectively by normalized kinetic
Reynolds stress, −u′w′/urmswrms, and quadratic windwise airflow velocities, u2) and windwise particle
displacement (Dx) to demonstrate rocking motions caused by (a) Q4 events (event A) and (b) Q1 events
(event B) alone. The subthreshold airflow condition is indicated by the average windwise velocity
18
Fig. 5. Representation of 1s histories of quadrant events (described respectively by normalized kinetic
Reynolds stress, −u′w′/urmswrms, and quadratic windwise airflow velocities, u2) and windwise particle
displacement (Dx) to demonstrate (a) rocking motions caused by a sequence of Q1 events following
by Q4 events (event C and event D, respectively) and (b) the incipient rolling caused by Q4 events.
The subthreshold airflow condition is indicated by the average windwise velocity below the threshold
19
Sweep events resulting in particle movements are characterized by randomness, and probabilistic
models can describe their features. Such as, the interval time between events follows a negatively
exponential distribution (Fig. 6a) which means the trend of intensive emergence of energetic sweep
𝑡 +𝑇𝑓
events. The duration of events Tf and event energy (~ ∫𝑡 0 𝑢3 𝑑𝑡 ) follow extreme value theory
0
distribution models (Fig. 6b and Fig. 6c, respectively) characterized by a right, long tail implying that
Fig. 6. Frequency distributions of (a) interval time Ti, (b) duration Tf and (c) energy content
𝑡 +𝑇𝑓
(~∫𝑡 0 𝑢3 𝑑𝑡) of sweep events resulting in particle movements.
0
Although sweep events are shown to be the relevant type of flow structures for particle incipient
motion, inspections of synchronous time histories of particle displacements and quadrant events
similar to Fig. 4 and Fig. 5 reveal that many sweep events do not lead to movements even though they
may induce high stress (e.g., event F in Fig. 4a, event G in Fig. 5a and event H in Fig. 5b). It appears
that only those events capable of inducing higher drag force than the initial surface resistance level
(parameterized by the fluid threshold, u2cr,0) are effective for particle movements. u2cr,0 is derived from
Bagnold’s force model using the mean drag coefficient Cd (=0.76) from Schmeeckle et al. (2007),
where spheres in near-bed turbulent flows were shown to have a larger drag coefficient than that for a
force (parameterized by the peak of quadratic windwise airflow velocities during the occurrence of
flow events, u2f,p) of most of the sweep events relevant to particle movements, more than 80%, is in
excess of u2cr,0. The data points below u2cr,0 may attribute to sweep events owning local drag
coefficients greater than the average value. Those events are sufficient to induce peak drag force
exceeding the local surface resistance level despite their relatively low peak flow velocities. The above
analyses support there is a force criterion (u2f,p>u2cr,0) for sweep events exciting rocking and rolling.
Fig. 7. Plot of the peak drag force (~u2f,p) against the duration (Tf) of sweep events resulting in particle
21
However, the force criterion u2f,p>u2cr,0 is no longer a sufficient condition for sweep events fully
dislodging particles. It is shown in Fig. 7 that the magnitude of peak drag force (~u2f,p) of sweep events
makes little sense in distinguishing incipient rolling from rocking motions. Thus, a force criterion alone
is insufficient to define full entrainment, as suggested by previous studies (Diplas et al., 2008;
The value of the critical vertical displacement zcr (=αd) for sweep events is acquired from the vertical
displacement measurements caused by sweep events (zf) in the ascending phase of rocking motions, as
shown in Fig. 8, which gives α = 0.053. Once the particle performs a greater vertical displacement, the
mean flow forcing, greater than the reduced resistance at the displaced location of greater exposure to
the flow, is enough to entrain the particle completely (similar to Valyrakis et al., 2011b).
22
Fig. 8. Plot of vertical displacements of particles resulting from sweep events (zf) in the ascending
phase of rocking motions against ascending duration (ta) of the particle showing the critical vertical
23
Fig. 9. Representation of 1s histories of the rate of work of air drag (~u2v) and energy density (~u3)
near the instance of rocking motions showing the calculation of normalized energy transfer coefficient
of sweep events, C′eff, which equals the work of air drag (area of grey shadow) divided by
corresponding event energy (area of stripe shadow). These examples also demonstrate the low value
of C′eff caused by relatively long sweep events where only a fraction of event energy is available to the
particle. Such as, the energy content of event A at the left of the vertical dashed line and event B at the
right of the vertical dashed line cannot be utilized by particles for downwind ascending motions.
The values of normalized energy transfer coefficient C'eff (=Ceff/Cd) defined by Eq. (6) are obtained
from synchronous time histories of energy density (~u3) and rate of work of drag force (~u2v). For
24
example, C'eff of sweep events shown in Fig. 9 (event A and event B) equals the work of air drag (the
area of grey shadows) divided by corresponding event energy (the area of stripe shadows). It is noted
that there may be multiple sweep events transferring energy to the same displacement events. In this
case, each event is assigned an individual value of C'eff. C'eff is distributed randomly following a
lognormal distribution model (Fig. 10a). However, a general trend is observed in which relatively
distinct C'eff values exist for rocking and incipient rolling (Fig. 10b). Generally, C'eff is from 0.0008 to
0.005 for rocking and from 0.005 to 0.018 for incipient rolling. This suggests that C'eff is an important
parameter to distinguish rocking from rolling. For sweep events entraining particles fully, C′eff should
Fig. 10. (a) Frequency distribution of normalized energy transfer coefficient, C′eff and (b) plot of event
𝑡 +𝑇𝑓 𝑡 +𝑇𝑓
energy (~∫𝑡 0 𝑢3 𝑑𝑡 ) against work of air drag (~∫𝑡 0 𝑢2 𝑣𝑑𝑡 ) of sweep events to demonstrate
0 0
25
The relationship between the magnitude of C'eff and structures of flow events is elusive. However,
if sweep events are long but only a fraction of event energy can be available for particle downwind
movements, C'eff is generally low. Two representative instances are provided in Fig. 9 (event A and
event B). In the first context (event A), the energy content in the head of events (flow energy on the
left of the vertical dashed line) cannot be utilized by particles because these flow components cannot
induce sufficient drag force exceeding the initial surface resistance level. In the second context (event
B), the energy content in the tail of events (flow energy on the right of the vertical dashed line) is not
available because these flow components cannot fight the temporal surface resistance level and then
Sweep events transferring energy to particles were extracted for analysis. If several sweep events
transfer energy to the same displacement events, the corresponding energy obtained by the particles is
the sum of energy offered by each event. Comparisons of threshold energy levels defined by Eq. (5) to
experimental measurements are presented in Fig. 11. The theoretical threshold level successfully
accounts for 94% of rolling and 99% of rocking motion. This result supports our hypothesis that sweep
26
′ 𝑡 +𝑇𝑓
Fig. 11. Plot of energy offered by sweep events (~𝐶𝑒𝑓𝑓 ∫𝑡
0
𝑢3 𝑑𝑡) to particles against ascending
0
duration of particles (ta). Parameter values in Eq. (5): ρp=80.8 kg/m3, d=40 mm, ρa=1.2 kg/m3, Cd=0.76,
g=9.8 m/s2 and α=0.053. The values of normalized energy transfer coefficient C'eff (=Ceff/Cd) defined
by Eq. (6) for each event are obtained from synchronous time histories of energy density (~u3) and rate
The energy offered by sweep events to particles (EfCeff) relates positively to the energy transfer
coefficient (Ceff) and duration of events (Tf). Thus, the combination of magnitudes of Ceff and Tf can
lead to various modes of achievement of full entrainment. For example, a long event with high Ceff
naturally entrains particles easily, similar to event A as shown in Fig 12. However, relatively short
27
events but owing high Ceff or long events with low Ceff (e.g., event B and event C in Fig. 12, respectively)
Fig. 12. Representation of 1s histories of the rate of work of air drag to demonstrate the single-event
mode of full entrainment (incipient rolling): (a) the full entrainment is caused by a relatively long
(Tf=0.196 s) sweep event (event A) with high energy transfer coefficient (C′eff =0.0069); (b) the full
entrainment is caused by a relatively short (Tf=0.088 s) sweep event (event B) and high energy transfer
coefficient (C′eff =0.0066); (c) the full entrainment is caused by a long sweep event (Tf =0.264 s) sweep
event (event C) and relatively low energy transfer coefficient (C′eff =0.0028). The vertical dashed line
Except for the single-event mode mentioned above, a portion of rolling motions (approximately
40%) is achieved by the combined actions of two successive sweep events, as shown in Fig. 13. In this
mode, the first sweep event suffices for setting a particle into motion; however, alone it cannot supply
sufficient energy to entrain the particle fully. Instead, it may push the particle to a an elevated location
of reduced resistance (e.g. see event A, in Fig. 13 a). Subsequently, the shortage of energy is
compensated by a trailing aerodynamic event (e.g. see event B, in Fig. 13 a). In some situations, the
28
following aerodynamic event, similar to event D in Fig. 13b, might be able to offer sufficient energy
for full entrainment on its own (e.g. not requiring the event C preciding).
Fig. 13. Representation of 1s histories of the rate of work of air drag to demonstrate the multi-event
mode for full entrainment (incipient rolling): (a) the full entrainment is caused by two sweep events
(event A and event B), and any of them alone cannot supply sufficient energy to particles; (b) the full
entrainment is caused by two sweep event (event C and event D), and the event D alone can supply
Considering how to cluster the impact of such aerodynamic events together, towards a single full
particle entrainment, can be seen as a methodological refinement that can be considered for future
analysis, along with considering other forcing contributions than direct aerodynamic drag. It is clear
that those events should have relative temporal proximity, so that their effect does not fade out with
the particle falling back into its initial resting pocket. At least, the time from the end of the previous
event to the onset of the following event should be shorter than the duration of ascending motions of
particles. For example, measurements indicate that the ratio of the time between two sweep events
29
transferring energy to the same displacement events to the corresponding ascending duration is from
0.11 to 0.41 for incipient rolling and from 0.08 to 0.52 for rocking. Another hint refers to the occurrence
of sweep events in airflow, which may inherently have a grouped pattern similar to water flow (Best,
1992).
5. Discussion
posing great challenges to the explanatory ability of the traditional viewpoint based on the mean flow.
On the other hand, the event-based flow viewpoint brings up a promising methodology to solve the
issue of intermittency because it isolates strong components from the turbulent flow by employing a
certain standard, and these isolated components may be the stockholders of local transport phenomena.
5.1. The utility of the quadrant method and the energy criterion
This study focuses on the role of coherent airflow events on the incipient motions of individual
gravel particles using the quadrant technique (Lu and Willmarth, 1973), which detected flow events
based on signs of fluctuation velocity and a threshold of Reynolds stress. The results suggest that the
incipient motions of gravel-size particles involving rocking and incipient rolling are attributed to
sweeps and outward interactions, both of which have positive windwise fluctuation velocities.
Furthermore, on flatbed surfaces, time histories of the windwise flow velocities instead of vertical flow
velocities or stress are coupled well with the time histories of the drag force (Schmeeckle et al., 2007).
This implies that only sweeps and outward interactions have the potential to induce a higher
instantaneous drag force than the initial surface resistance level under subthreshold flow conditions
where the time-averaged drag force is lower than the initial surface resistance level.
30
The quadrant technique provides a rough empirical approach for identifying flow events
responsible for particles' incipient entrainment. It classifies high-stress events into four types, with
clearly sweeps contributing the majority of cases leading to coarse particles' incipient entrainment, as
shown in Table 2. However, quadrant analysis only considers the magnitude of aerodynamic forces
thus being unable to establish which amongst the sweeps would be effective to fully dislodge particles
downwind, as demonstrated in Fig. 4 and Fig. 5. This study similar to Valyrakis et al. (2013), considers
that only quadrant events of sufficient aerodynamic force and energy intensity can lead to complete
aerodynamic entrainment. The force criterion ensures that sweep events can set particles into motion.
However, alone it cannot determine whether the resulting motion is a rocking event (incomplete
entrainment) or an incipient rolling (full entrainment) because the incipient rolling occurs only when
a critical vertical displacement is achieved. The presented energy criterion links the critical vertical
displacement to the energy content of the quadrant events, required to perform the critical mechanical
work. It synthesizes the features of sweep events (e.g., energy transfer coefficient Ceff and duration Tf),
both having influences on incipient rolling (full entrainment) of particles and successfully
The theoretical frame of particle entrainment has seen innovative developments in recent years
(e.g., Diplas et al., 2008; Celik et al., 2010; Valyrakis et al., 2010; Valyrakis et al., 2011b; Valyrakis et
al., 2013). These frameworks are based on the transfer of flow energy or momentum and are distinct
from the traditionally employed Shields or Bagnold's criteria for identifying the flow conditions for
intermittent entrainment of particles in turbulent flows. Such flow events are called energetic flow
events (Valyrakis et al., 2013), similar to flow components above the dash-dot line (u3cr,0) in Fig. 9.
They are characterized by all components of the flow events capable of inducing sufficient drag force
31
in excess of the surface resistance level. It may be interesting to explore the relationship between
energetic flow events and sweep events, considering both can entrain particles. As visualized in Fig. 9,
a simple analysis reveals that the energetic flow and sweep events resulting in particle movements are
partially overlapping. However, the definition of sweep events only implies a flow stress criterion and
lacks any connection to particle properties. Therefore, the force criterion developed using the windwise
flow velocity should be presented to help distinguish which sweep events can induce sufficient drag
force to exceed the initial surface resistance level. In addition, energetic sweep events can involve
ineffective aerodynamic forcing components that do not transfer energy to particles, as event A and
event B are shown in Fig. 9. Considering the above, the energetic flow events criterion is superior to
the sweep events method in assessing incipient particle entrainment because it has embeded the initial
aerodynamic forcing.
The rocking and rolling of gravel particles are relatively new topics for aeolian transport and
amongst foundational dynamic processes for forming and evolving aeolian coarse-gravel mega-ripples.
Specifically, both movements can be initiated aerodynamically without requiring the impact of fine
materials such as sand grains (de Silva et al., 2013). However, previous numerical models for
mega-ripples (Yizhaq, 2004; Yizhaq, 2008) generally neglect these aerodynamic processes, consider
the entrainment of coarse particles by the impact of sand grains, and naturally cannot provide
comprehensive descriptions for mega-ripples. This study is the first attempt to investigate the
interaction between coarse particles and turbulent airflow. The presented force and energy criteria
could be used to estimate the occurrence probability (P) of intermittent rocking and rolling in an
32
appropriate probabilistic framework (e.g., extreme value theory suggested by Valyrakis et al., 2011b),
considering both Ef and C′eff are random variables and large value of them conducive to full entrainment
are rare events implying by their long-tail probability distribution modes (see Fig. 6c and Fig. 10a).
For example, the rocking probability can be expressed as the product of P(u2f,p>u2cr) and P(EfCeff<Wp,cr).
Similarly, the rolling probability is the product of P(u2f,p>u2cr) and P(EfCeff >Wp,cr). If the number
density of sweep events in a specific region has been obtained, the number density of fully entrained
particles in this region, which may be necessary for modeling mega-ripples, is the product of rolling
probability and number density of sweep events. However, this idea is only available for single-event
entrainment mode. For the multi-event entrainment mode, the idea may be available only if merging
multiple adjacent sweep events and transferring energy to the same displacement events has been
further developed.
6. Conclusions
The intermittent creep of gravel particles, as well as particle rocking, is of importance for assessing
geomorphologic work and estimating aeolian transport processes. This study develops a micro-
-mechanical model to characterize the incipient motion of gravel particles assumed to be initiated by
coherent airflow structures (quadrant events). A work-based criterion Eq. (5), is established from the
The proposed criterion agrees well with observations obtained from wind tunnel experiments. On
the flat and rough bed surface, it was found that both rocking and rolling are initiated primarily by
sweep events capable of inducing sufficient drag force to exceed the surface resistance. Beyond that
instance, rolling will occur if the aerodynamic forcing event has succificent duration and energy
33
intensity for fully dislodging the rolling particle. Otherwise the particle’s motion is classified as a
rocking event, where the particle returns back to its resting pocket. These latter type of motions do not
contribute to aeolian transport, other than indirectly, as is done for example in the case they are enabling
mechanical sieving. The proposed criterion can be used for indirectly estimating natural coarse (such
as sand or gravel) particles' rocking and rolling probabilities near threshold flow conditions, as well as
directly, for the assessment of dynamic thresholds of anthropogenic pollutants of similar shape and
Acknowledgments: This research has been supported by the National Natural Science Foundation of
China (Grants No. 41171005, 41071005, 12272344) and the Ministry of Science and Technology of
the People's Republic of China (Grant No. 2013CB956000).
Open Research
Data availability statement: The data (Valyrakis and Xiao-Hu, 2024) used in this manuscript, as well
as for the wind velocity profile and turbulence spectra are archived in a repository, and can be accessed
via the following URLs: https://zenodo.org/records/7782901.
References
Adrian, R.J., 2007. Hairpin vortex organization in wall turbulence. Physics of Fluids 19, 041301.
Anderson, R.S., 1987. Eolian sediment transport as a stochastic process: The effects of a fluctuating
Baas, A.C.W., Sherman, D.J., 2005. Formation and behavior of aeolian streamers. Journal of
34
Batt, R.G., Peabody, S.A., 1999. Threshold friction velocities for large pebble gravel beds. Journal of
Bauer, B.O., Walker, I.J., Baas, A.C.W., Jackson, D.W.T., Neuman, C.M., Wiggs, G.F.S., Hesp, P.A.,
2013. Critical reflections on the coherent flow structures paradigm in aeolian geomorphology.
In: Venditti J. G., Best, J. L., Church, M., Hardy, R. J. (Eds.), Coherent Flow Structures at Earth's
Bauer, B.O., Yi, J.C., Namikas, S.L., Sherman, D.J., 1998. Event detection and conditional averaging
Best, J., 1992. On the entrainment of sediment and initiation of bed defects: insights from recent
Bogard, D.G., Tiederman, W.G., 1986. Burst detection with single-point velocity measurements.
Burton, T., Jenkins, N., Sharpe, D., Bossanyi, E., 2011. Wind energy handbook. John Wiley & Sons
Chapman, C.A., Walker, I.J., Hesp, P.A., Bauer, B.O., Davidson-Arnott R.G.D., 2012. Turbulent
Reynolds stress and quadrant event activity in wind flow over a coastal foredune.
Celik, A.O., Diplas, P., Dancey, C.L., Valyrakis, M., 2010. Impulse and particle dislodgement under
de Silva, S.L., Spagnuolo, M.G., Bridges, N.T., Zimbelman, J.R., 2013. Gravel-mantled megaripples
of the Argentinean Puna: A model for their origin and growth with implications for Mars.
35
Detert, M., Nikora, V., Jirka, G.H., 2010. Synoptic velocity and pressure fields at the water-sediment
Diplas, P., Dancey, C.L., Celik, A.O., Valyrakis, M., Greer, K., Akar, T., 2008. The role of
Diplas, P., Celik, A.O., Dancey, C.L., Valyrakis, M. 2010. Non-intrusive method for Detecting Particle
Movement Characteristics near Threshold Flow Conditions, Journal of Irrigation and Drainage
Drake, T.G., Shreve, R.L., Dietrich, W.E., Whiting, P.J., Leopold, L.B., 1988. Bedload transport of fine
gravel observed by motion picture photography. Journal of Fluid Mechanics 192, 193−217.
Durán, O., Claudin, P., Andreotti, B., 2011. On aeolian transport: Grain-scale interactions, dynamical
Dwivedi, A., Melville, B.W., Shamseldin, A.Y., Guha, T.K., 2011. Flow structures and hydrodynamic
Gillies, J.A., Nickling, W.G., Tilson, M., Furtak-Cole, F., 2012. Wind-formed gravel bed forms, Wright
Grass, A.J., 1971. Structural features of turbulent flow over smooth and rough boundaries, Journal of
Greeley, R., Iversen, J.D., 1985. Wind as a Geological Process on Earth, Mars, Venus, and Titan.
Hardy, R.J., Best, J.L., Lane, S.N., Carbonneau, P.E., 2009. Coherent flow structures in a depth-limited
flow over a gravel surface: The role of near-bed turbulence and influence of Reynolds number.
36
Journal of Geophysical Research 114, F01003.
Hofland, B., Booij, R., 2004. Measuring the flow structures that initiate stone movement. In: Greco,
M., Carravetta, A., Della Morte, R., Balkema, A.A. (Eds. ), River Flow 2004. Rotterdam,
Isenberg, O., Yizhaq, H., Tsoar, H., Wenkart, R., Karnieli, A., Kok, J.F., Katra, I., 2011. Megaripple
Jackson, R.G., 1976. Sedimentological and fluid-dynamic implications of the turbulent bursting
Kaftori, D., Hetsroni, G., Banerjee, S.,1995. Particle behavior in the turbulent boundary layer. II.
Katra, I., Yizhaq, H., Kok, J.F., 2014. Mechanisms limiting the growth of aeolian megaripples.
Keylock, C.J., Lane, S. N., Richards, K. S., 2014. Quadrant/octant sequencing and the role of coherent
structures in bed load sediment entrainment. Journal of Geophysical Research: Earth Surface
119, 264−286.
Kline, S.J., Reynolds, W.C., Schraub, F.A., Runstadler, P.W., 1967. The structure of turbulent boundary
Kok, J.F., Renno, N.O., 2009. A comprehensive numerical model of steady-state saltation(COMSALT).
Leenders, J.K., van Boxel, J.H., Sterk, G., 2005. Wind forces and related saltation transport.
Livingstone, I., Wiggs, G.F.S., Weaver, C.M., 2007. Geomorphology of desert sand dunes: A review
37
of recent progress. Earth-Science Reviews 80, 239–257.
Li, Z.S., Zhao, X.H., Huang, W., 2008. A stochastic model for initial movement of sand grains by wind.
Lu, H., Raupach, M.R., Richards, K.S., 2005. Modeling entrainment of sedimentary particles by wind
Lu, S.S., Willmarth, W.W., 1973. Measurements of the structure of the Reynolds stress in a turbulent
Mazumder, R., 2000. Turbulence-particle interactions and their implications for sediment transport and
Milana, J. P., 2009. Largest wind ripples on Earth? Geology 37, 343−346.
Pähtz, T., Clark, A. H., Valyrakis, M., Durán, O., 2020. The physics of sediment transport initiation,
cessation, and entrainment across aeolian and fluvial environments. Reviews of Geophysics 58,
e2019RG000679.
Pähtz, T., Valyrakis, M., Zhao, X. H., Li, Z. S., 2018. The critical role of the boundary layer thickness
Pöschel, T., Schwager, T., Brilliantov, N. V., 1999. Rolling friction of a hard cylinder on a viscous
Robinson, S.K., 1991. Coherent Motions in the turbulent boundary layer. Annual Review of Fluid
Schmeeckle, M.W., Nelson, J.M., Shreve, R.L., 2007. Forces on stationary particles in near-bed
38
Schönfeldt, H.J., von Löwis, S., 2003. Turbulence-driven saltation in the atmospheric surface layer.
Shao, Y.P., Lu H., 2000. A simple expression for wind erosion threshold friction velocity. Journal of
Spies, P., McEwan I.K., 2000. Equilibration of saltation. Earth Surface Processes and Landforms 25,
437−453.
Sterk, G., Jacobs, A.F.G., van Boxel, J.H., 1998. The effect of turbulent flow structures on saltation
sand transport in the atmospheric boundary layer. Earth Surface Processes and Landforms 23,
877−887.
Tholen, K., Pähtz, T., Yizhaq, H., Katra, I., Kroy, K., 2022. Megaripple mechanics: bimodal transport
Thorne, P.D., Williams, J.J., Heathershaw, A.D., 1989. In situ acoustic measurements of marine gravel
Valyrakis, M., Diplas, P., Dancey, C.L., Greer, K., Celik, A.O., 2010. Role of instantaneous force
115, F02006.
Valyrakis M., Diplas, P., Dancey, C.L., 2011a. Analysis of impulse events associated to entrainment of
coarse particles. Coherent Flow Structures in Geophysical Flows at Earth Surface. Simon
Valyrakis, M., Diplas, P., Dancey, C.L., 2011b. Entrainment of coarse grains in turbulent flows: An
extreme value theory approach. Water Resources Research, 47, W09512, DOI:
39
10.1029/2010WR010236.
Valyrakis, M., Diplas, P., Dancey, C.L., 2013. Entrainment of coarse particles in turbulent flows: An
Valyrakis, M., Z. Xiao-Hu (2024). Wind Data for the Mean Velocity Profile and Turbulence Spectra
van Boxel, J.H., Sterk, G., Arens, S.M., 2004. Sonic anemometers in aeolian sediment transport
Walker, I.J., 2005. Physical and logistical considerations of using ultrasonic anemometers in aeolian
Wiggs, G.F.S., Weaver, C.M., 2012. Turbulent flow structures and aeolian sediment transport over a
Wu, F.C., Shih, W.R., 2012. Entrainment of sediment particles by retrograde vortices: Test of
Xiao-Hu, Zh., Valyrakis, M., Zhen-Shan L. 2021. Rock and roll: incipient aeolian entrainment of
Yager, E. M., Schmeeckle, M. W., Badoux, A. 2018. Resistance is not futile: Grain resistance
controls on observed critical Shields stress variations. Journal of Geophysical Research: Earth
Yizhaq, H., 2004. A simple model of aeolian megaripples. Physica A 338, 211−217.
Yizhaq, H., 2008. Aeolian megaripples: Mathematical model and numerical Simulations. Journal of
Yizhaq, H., Katra, I., Isenberg, O., and Tsoar, H., 2012. Evolution of megaripples from a flat bed.
40
Aeolian Research 6, 1−12.
Zimbelman J.R., Irwin III, R.P., Williams, S.H., Bunch, F., Valdez, A., Steven, S., 2009. The rate of
41