You are on page 1of 37

J. Fluid Mech. (2013), vol. 733, pp. 588624.

doi:10.1017/jfm.2013.435

c Cambridge University Press 2013


588

Airwater gas transfer and near-surface motions


Damon E. Turney and Sanjoy Banerjee
Department of Chemical Engineering, The City College of New York, NY 10031, USA
(Received 31 October 2012; revised 3 June 2013; accepted 14 August 2013;
first published online 26 September 2013)

Rates of gas transfer between air and water remain difficult to predict or simulate due
to the wide range of length and time scales and lack of experimental observations
of near-surface fluid velocity and gas concentrations. The surface renewal model (SR)
and surface divergence model (SD) provide the two leading models of the process,
yet they remain poorly tested by observation because near-surface velocity is difficult
to measure. To contribute to evaluation of these models, we apply new techniques
called interfacial particle imaging velocimetry (IPIV) and three-dimensional IPIV (3DIPIV) for measuring water velocities within a millimetre of a moving deformable
airwater interface. The latter technique (3D-IPIV) simultaneously measures the
airwater interface topography. We apply these techniques to turbulent open-channel
water flows and wind-sheared water flows with microscale breaking waves. Additional
measurements made for each flow condition are bulk turbulent length scales, bulk
turbulent velocity scales, airwater gas transfer rates, friction velocities, and wave
characteristics. We analyse these data to test the surface divergence models for
interfacial gas transfer. The first test is of predictions from the Banerjee (Ninth
International Heat Transfer Conference, Keynote Lectures, vol. 1, 1990, pp. 395418,
Hemisphere Press) surface divergence model for gas transfer for homogeneous
isotropic turbulence interacting with a planar free surface. The second test is of
predictions from the McCready, Vassiliadou and Hanratty (AIChE J., vol. 32(7),
1986, pp. 11081115) surface divergence model, as applied in both open-channel
flow and wind-sheared wavy flows. We find the predictions of the Banerjee and
McCready et al. models to agree with the experimental data taken for open-channel
flow conditions. On the other hand, for wind-driven flows with wind waves we find
disagreement between the McCready et al. predictions and our direct measurements
of the gas transfer coefficient. The cause of the disagreement is investigated by
Lagrangian tracking of surface divergence of surface water patches, and by analysis
of the corresponding Lagrangian time series with advectiondiffusion concepts. A
quantitative criterion based on surface divergence strength and lifetime is proposed
to distinguish the effectiveness of each near-surface motion toward causing interfacial
gas transfer. Capillary waves are found to contribute to surface divergence but to
have too short a time scale to cause interfacial gas transfer. As wind speed increases,
the presence and intensity on the airwater interface of capillary waves and other
ineffective near-surface motions is diminished by the rise of turbulent wakes from
microscale breaking waves thus causing the disagreement of the surface divergence
models predicted transfer rates with measurements. A model of airwater gas transfer
that combines the surface renewal and surface divergence models is formulated and

Email address for correspondence: dturney@ccny.cuny.edu

Airwater gas transfer and near-surface motions

589

found to agree with the data from both open-channel flows and wind-driven flows
without requiring an empirical coefficient.
Key words: gas/liquid flows, geophysical and geological flows, waves/free-surface flows

1. Introduction

Transfer of sparingly soluble gas species across gasliquid interfaces is important


for a wide variety of systems, e.g. in industrial equipment such as scrubbers,
absorbers and gasliquid reactors, as well as in environmental processes such as
oceanic absorption of atmospheric CO2 and uptake of mercury emissions from coal
power plants. Interfacial gas transfer is also analogous to interfacial heat transfer
in condensers and evaporators. Direct use of the advectiondiffusion equation for
prediction of the transfer rate is rarely feasible because the turbulent and otherwise
complex near-interface fluid motions are difficult to measure or obtain from simulation.
Simplifying models have been developed to allow prediction of transfer rates, but
the typical uncertainty in these predictions is large and ranges between a factor
of 2 and 10 (Belanger & Korzun 1991; Dvorak et al. 1996; Piche, Grandjean
& Larachi 2002; Wanninkhof et al. 2004; Ho et al. 2006). This paper reports
laboratory measurements of near-surface motions and interfacial gas transfer rates
in airwater systems in turbulent open-channel flows and wind-driven flows including
microscale wave breaking. The data are used to elucidate the important near-surface
flow phenomena to guide the development and testing of predictive models.
The scope of this work focuses on airwater systems where the main resistance
to interfacial transfer of gases lies on the water side of the interface. As mentioned
above, many systems of interest fall within this category (Jahne & Hauecker 1998;
Turney 2009), where air-side motions have little effect on the interfacial transfer
rate because of the greater diffusivity of the gases in air compared with water, and
low gas solubility in water. Both these factors allow the air-side gas concentration
to homogenize while a concentration boundary layer forms in the water along the
interface and interacts with near-surface motions. The advectiondiffusion equations
describe the timespace evolution of the concentration field:
c
= F,
t
F = vc Dc,

(1.1)
(1.2)

where c is the gas concentration, F is the flux of gas species, v is fluid velocity, and
D is the gas diffusivity in water. The NavierStokes equations govern the velocity field.
The instantaneous gas flux across the airwater interface Fi is by diffusion:

c
Fi = D ,
n i

(1.3)

where the subscript i denotes a location on the airwater interface and n is the distance
normal to the interface. Since these governing equations are linear with respect to
concentration, the time-averaged flux across the interface must be proportional to the

590

D. E. Turney and S. Banerjee

concentration difference at the system boundaries, e.g. the bulk fluid, and we can write

c
(1.4)
F i = k(Scba cbw ) = D ,
n i
where the overbars denote time or space averaging, cba is the gas concentration in the
bulk air, S is the gas equilibrium solubility in water, cbw is the concentration in the
bulk water and k is the time-averaged gas transfer coefficient. The value of the transfer
coefficient k depends on the gas diffusivity in water and the water motions that
affect the thin gas concentration gradient at the interface. The governing equations are
difficult to solve computationally, therefore k is usually estimated by simplified models
such as surface renewal theory or surface divergence theory. The most commonly used
models are listed in table 1 and briefly explained in the following paragraph.
The earliest models of the turbulent process (the film model of Lewis & Whitman
1924) predicted that mass transfer rates would be proportional to D, which was not
supported by experiments. In response to this, Higbie (1935) modelled mass transfer
by calculating the time-dependent mass diffusion rate into a liquid layer exposed to
a sudden change in concentration, which gives a proportionality with the square root
of D. Danckwerts (1951) included an integration of Higbies model with a stochastic
description of the turbulent near-surface motions, assuming that each location on
the gasliquid interface had aged a Poisson-distributed length of time since it
had been renewed with bulk liquid. Danckwerts assumed that each renewal event
instantaneously homogenized the near-surface liquids gas concentration to cbw , and
that the liquid was stagnant until the next renewal. Danckwerts integrated Higbies
equation over all Poisson-distributed values of surface age, producing the surface
renewal (SR) model shown in (1.7). The only unknown was the mean time between
the renewal events in the Poisson distribution, a parameter called the mean surface
renewal time . Fortescue & Pearson (1967) assumed this surface renewal time to be
equal to the mean turnover time of the largest eddies of a turbulent flow L/u0 .
This assumption leads to the large-eddy SR model given by (1.8). Fortescue & Pearson
supported this model with some limited experimental data. At about the same time
Banerjee et al. (1968) and Lamont & Scott (1970) proposed a model in which the
viscous dissipation range small-scale eddies control the renewal motions in the SR
model. These small-scale
eddies have a mean turnover time scaled by Kolmogorov

parameters /, leading to the small-eddy SR model which was also supported


by limited experimental data. Each of these models gave quite different dependence
on the turbulent Reynolds numbers, as seen in (1.8) and (1.9). Theofanous, Houze &
Brumfield (1976) provided an explanation that resolved the discrepancy between the
large- and small-eddy SR models by noticing that Fortescue & Pearsons experiments
were at low turbulent Reynolds numbers for which mass transfer is dominated by large
eddy sizes, and that those of Banerjee et al. and Lamont & Scott were at higher
turbulent Reynolds numbers for which mass transfer should be governed by small
eddies.
In subsequent years many attempts were made to directly measure the mean time
between surface renewals rather than estimate it with bulk turbulence parameters
(Komori, Yasuhiro & Hiromasa 1989; Rashidi & Banerjee 1990a,b), but it was difficult
to unambiguously identify renewal events. For example, upwellings in open-channel
flows could be considered renewal events, but not all appeared to qualify. Komori
et al. (1989) mention this problem when interpreting their open-channel gas absorption
experiments and needed to use a corrective coefficient to fit their measured gas
transfer rate to the SR model. Unfortunately the corrective coefficient varied between

film
penetration
surface renewal (SR)
large-eddy SR
small-eddy SR
SDM
SDB

Model name

k
= D/
k =
D/( )
= D/
k

k = D/ ,
L/u0 giving k = cu0 Sc1/2 Re1/2
t
k = D/ , / giving k = cu0 Sc1/2 Re1/4
t
0
k = c D

1/4
k = cu0 Sc1/2 Re1/2
0.3(2.83Re3/4
2.14Re2/3
t
t
t

Calculation of transfer coefficient


(1.5)
(1.6)
(1.7)
(1.8)
(1.9)
(1.10)
(1.11)

TABLE 1. A list of the models of airwater gas transfer. For these equations D is the chemical diffusivity, is the LewisWhitman thickness
of the concentration boundary layer, u0 is the r.m.s. velocity fluctuation in the bulk flow, is the mean time between surface renewal events,
L is the bulk turbulence integral length scale, Sc is the Schmidt number, Ret is a turbulent Reynolds number (i.e. Ret = u0L L/v where u0L is
the turbulence intensity at a distance L from the interface), c values are empirically determined coefficients, 0 is the r.m.s. of the divergence
of the (two-dimensional) interfacial velocity field, SR stands for surface renewal, SDM stands for the surface divergence model of McCready,
Vassiliadou & Hanratty (1986) and SDB stands for the surface divergence model of Banerjee (1990).

Lewis & Whitman (1924)


Higbie (1935)
Danckwerts (1951)
Fortescue & Pearson (1967)
Banerjee, Scott & Rhodes (1968)
McCready et al. (1986)
Banerjee (1990)

Authors

Airwater gas transfer and near-surface motions


591

592

D. E. Turney and S. Banerjee

experiments and flow configurations, and it became difficult to predict mass transfer
using the SR model.
A different approach was proposed by McCready et al. (1986) by modelling the
interfacial transfer process in a manner similar to Chan & Scriven (1970). The
approach uses the governing equations (advectiondiffusion equation) to suggest that
the gasliquid transfer rate is controlled by surface-normal motions that each have the
form of a stagnation flow. The key simplifications are that at high Sc the interfaceparallel motions have a negligible effect compared to the interface-normal motions,
and that only near-surface (viscous boundary layer) motions need consideration.
Therefore the velocity field used with the advectiondiffusion equation is the first
term in a Taylor series expansion,


ui wi
y = y,
(1.12)
+
vi =
x
z
where ui , vi and wi are the near-surface velocities in the x, y, and z directions (y
being oriented normal to the local interface with y = 0 at the interface). The terms in
parenthesis are the surface divergence strength and are labelled for convenience.
Thus the model is called the surface divergence (SD) model. Chan & Scriven (who
included a term for a curved interface) found an analytical solution for mass transfer
for this form of the interface normal velocity field. McCready et al. (1986) built
on this and considered two limiting scenarios for predicting the interfacial transfer
rate, one for high frequency sinusoidal oscillations in and one for low frequency
oscillations. Their criterion for distinguishing between high or low frequency was
2
D/lcbl
> , where lcbl is the thickness of the concentration boundary layer and is
the frequency of a surface divergence oscillation. Most uses of McCready et al.s
formulation assume the low-frequency version, which is given in (1.10) in table 1.
This model (which we will call the SDM) identifies the root mean square (r.m.s.)
divergence 0 as the central parameter, and has been extensively used for airwater
gas transfer predictions (Tamburrino & Gulliver 2002; Banerjee & Macintyre 2004;
McKenna & McGillis 2004; Turney, Smith & Banerjee 2005; Magnaudet & Calmet
2006; Xu & Khoo 2006; Herlina & Jirka 2008; Turney & Banerjee 2008). The
advantages of the surface divergence model as compared to the surface renewal model
is elimination of the ambiguous task of defining the mean time between renewals ,
replacing it by measurements of the divergence of the surface velocities , which
though difficult is unambiguous.
The surface divergence information needed as input for the SD model can be
estimated theoretically for simple cases. This approach was taken by Brumley & Jirka
(1987), using a theory of homogeneous isotropic turbulence interacting with a clean
free surface proposed by Hunt & Graham (1978). Banerjee (1990) then proposed a
model combining aspects of the results found by Hunt & Graham (1978), McCready
et al. (1986), and Brumley & Jirka (1987) to calculate the interfacial mass transfer
coefficient for homogeneous isotropic turbulence impacting on a free surface, given as
(1.11) in table 1. As detailed in Banerjee (1990), the constants in (1.11) arise from a
fit to the homogeneous isotropic turbulence energy spectrum at distances far from the
surface. Strictly the model should be applied only to cases where there is no shear
at the interface, but it has been used even for cases with high shear (see Banerjee
& Macintyre 2004) as it only requires an estimate of the turbulent Reynolds number
based on integral scales far from the interface. This model (which we will call SDB)
is a variant of the SD model and agrees with direct numerical simulations (Banerjee &
Macintyre 2004) and some field data (Banerjee & Macintyre 2004).

593

Airwater gas transfer and near-surface motions


Publication
McCready et al. (1986)
Law & Khoo (2002)
Tamburrino & Gulliver (2002)
McKenna & McGillis (2004)
Banerjee & Macintyre (2004)
Turney et al. (2005)
Magnaudet & Calmet (2006)
Xu & Khoo (2006)
Herlina & Jirka (2008)

Flow condition
counter-current wind
shear
grid-stirred tank and
wind waves
open-channel flow
grid-stirred tank
low wind flow
low and moderate wind
flow
LES of channel flow;
single condition
grid-stirred tank and
wind waves
grid-stirred tank

Calculation of transfer
coefficient

k 0.71 D 0

k 0.22 D 0
p
k 0.24 DSmax /ucb

k 0.50D 0
k 0.35D 0
k 0.50 D 0

k 0.60 D 0

k 0.20 D 0

k 0.33 D 0

TABLE 2. Previous tests of the McCready et al. (1986) surface divergence theory.

The SDB and SDM models have received some support but remain to be rigorously
tested by direct measurements of the surface divergence field, bulk turbulence
parameters, and mass transfer coefficients. This is partly because of the difficulty
of measuring near-surface velocity exclusively within 1 mm of the interface. When
wind waves cause the interface to rapidly deform, the difficulty becomes greater still.
Table 2 shows some published tests of the SDM model. Variations in the coefficient
of proportionality greater than a factor of three are seen. Each of these tests is now
discussed.
McCready et al. (1986) avoided the difficulty of making near-surface velocity
measurements at the airwater interface by using velocity measurements near the
solid wall of pipe flow from previous reports. They used a time series of pipe flow
near-wall velocity measurements to estimate time series for a finite difference
simulation of the advectiondiffusion equation. Their results produced the top equation
in table 2. The early methods for direct measurement of near-surface motion at
gasliquid interfaces used flow tracer particles that stay trapped on the interface
(Kumar & Banerjee 1998; Tamburrino & Gulliver 2002, 2007; Turney et al. 2005;
Tsumori & Sugihara 2007). The particle image velocimetry (PIV) data resulting
from these investigations suffered from flow tracers colliding with each other at
convergent downwellings and becoming too sparse for PIV in divergent upwellings.
Other investigations overcame this problem by mixing the flow tracers homogeneously
throughout the water and setting up the PIV camera to preferentially see tracers
near the interface rather than away from it (Law, Khoo & Chu 1999; Law & Khoo
2002; Peirson & Banner 2003; McKenna & McGillis 2004; Siddiqui et al. 2004;
Siddiqui & Loewen 2007; Xu & Khoo 2006; Herlina & Jirka 2008). Peirson & Banner
(2003) made side-view PIV measurements within 1 mm of the airwater interface
in wind-driven flows with microscale breaking waves, and reported a few data points
for velocities and flow divergences at the interface. Their measurements will be
compared to our data in 4. Peirson & Banner made no gas transfer measurements
and so could not directly test any interfacial transfer models. The data contained in
Law et al. (1999), Law & Khoo (2002) and Xu & Khoo (2006) also used side-view

594

D. E. Turney and S. Banerjee

PIV with a wavy interface identification technique. Their measurements of and k


came from a flume in the shape of an annulus 45 cm in diameter. Their analysis found
confirmation of the SDM model as shown in table 2. However, the small diameter
of their annulus flume and other anomalies must have strongly affected their surface
divergence dataset. Their findings will also be compared with our observations in
4. McKenna & McGillis (2004) made measurements of using an overhead PIV
camera looking down on a grid-stirred tank. They used a camera system with a depth
of focus of 510 mm, and reported their camera system to see only the velocities
within 5 mm or less of the airwater interface. Their measurements also supported the
SDM model, yet it is desirable to measure the velocities closer to the interface than
5 mm. Siddiqui et al. (2004) and Siddiqui & Loewen (2007) report side-view PIV
simultaneously with plan-view infrared imagery in wind-driven flows. Their results
provide insight into the production of turbulence and vorticity by microscale breaking
wind waves, but again no measurements of or k were made. Herlina & Jirka
(2008) used a grid-stirred tank to take side-view PIV measurements simultaneously
with laser-induced fluorescence measurements of the gas concentration field. These
data allowed Herlina & Jirka to calculate gas transfer rates, 0 values, and turbulence
statistics. Their conclusions supported unification of the small-eddy SR model at high
turbulent Reynolds numbers and large-eddy SR models at low turbulent Reynolds
numbers. Additionally their data supported the SDM model with a coefficient of
0.33 as shown in table 2. An outlier equation in table 2 is Tamburrino & Gulliver
(2002), who measured power spectra of in wavelength space. After conversion into
frequency space using a Taylors frozen turbulence assumption, they determined that
most of the power spectra content was at frequencies higher than the critical frequency,
so they used a different form of McCready et al.s surface divergence model. Although
valuable, the near-surface velocity data of Tamburrino and Gulliver used surface-bound
flow tracers and are probably affected by the problems mentioned earlier regarding
surface-bound particles.
Direct numerical simulations have also been utilized to test airwater gas transfer
models (Banerjee, Lakehal & Fulgosi 2004; Magnaudet & Calmet 2006). Banerjee
et al. (2004) provided confirmation of the SDB model at low wind speeds
U10 < 1.5 m s1 , where U10 is the estimated wind speed at 10 m height, resulting
in a coefficient of 0.35 as shown in the first row of table 3. The wind speeds in these
simulations were too low to produce microscale breaking wind waves. Simulation
at higher wind speed has so far proved too computationally expensive. Banerjee
et al. (2004) also tested the SDB model using experimental data from the grid-stirred
tank experiments of Chu & Jirka (1992) and McKenna & McGillis (2004), showing
agreement with a coefficient of 0.20 but with significant scatter in the fit. Magnaudet
& Calmet (2006) made large-eddy numerical simulations of channel flow and tested
the SDM model for a single flow condition, finding a coefficient of 0.6 which is
greater than most other tests, as shown in table 2.
This current study fits into the data gap in the existing literature by providing
plan-view near-surface velocity measurements (within 1 mm of a wavy airwater
interface) more free from PIV anomalies due to surface-bound flow tracers, together
with direct measurements of the airwater gas transfer rate and bulk turbulence
parameters. The SDM and SDB models are tested in open-channel flows and windy
conditions including microscale wave breaking. The data reported also provide insight
into the spatial and temporal scales of the divergence of the surface velocity field, and
ultimately motivate a re-examination of the applicability of the SD model, suggesting a
need for a time scale based on ideas similar to those in the SR model.

grid-stirred tank
low wind-shear; no waves

Flow condition

1/4
k = 0.20u0 Sc1/2 Re1/2
[0.3(2.83Re3/4
2.14Re2/3
t
t
t )]
1/4
k = 0.35u0 Sc1/2 Re1/2
[0.3(2.83Re3/4
2.14Re2/3
t
t
t )]

Calculation of transfer coefficient

TABLE 3. Previously published tests of the Banerjee (1990) surface divergence model SDB.

Banerjee et al. (2004)


Banerjee et al. (2004)

Publication

Airwater gas transfer and near-surface motions


595

596

D. E. Turney and S. Banerjee


Traverse support beam
All data collected
in this region

Light source
and cameras

Flow
straightener

Exhaust

Fan in

Glass window section


Water
pump
Water flow

F IGURE 1. Schematic of the experimental facility. The upstream section held gas exchange
equipment to remove all oxygen from the upstream water. High-speed cameras were mounted
on an overhead plan-view robotic traverse, or from side-view angles looking through glass
walls. A skimmer pump in the downstream tank continuously removed the top millimetre of
water and discarded it out of the system.

2. Experiments
All experimental measurements were conducted at the University of California,
Santa Barbara, in a horizontal channel of length 12.0 m, width 0.70 m, and height
0.30 m through which a stratified airwater channel flow or wind-driven flow was
established. A schematic of the experimental facilities is shown in figure 1. The ratio
of channel width to water depth was kept above the 5 criterion known to keep
sidewall secondary circulation away from the central region of the channel (Nezu
& Nakagawa 1993). Likewise the ratio of streamwise location of measurements to
water depth was kept above 60 to ensure that entrance effects were eliminated and
turbulence was fully developed, as is confirmed for such ratios above 40 by Nezu
& Nakagawa (1993). All air and water flows were concurrent. The upstream end of
the channel was equipped with flow straighteners, a wave suppressor, dissolved oxygen
stripping capability, and a large fan for generating wind inside the channel. The final
few metres of the channel held high-speed side-view PIV cameras, an overhead (planview) set of high-speed PIV cameras mounted on a 3 m long traverse to allow them
to travel downstream, a series of 3 mm diameter channel-bottom flush-mounted ports
for measuring dissolved oxygen concentration, capacitance wire wave-height gauges,
and Pitot tube wind speed sensors. The downstream end of the channel held wave
dampening structures, thermometers, and a surface water vacuum that continuously
skimmed and discarded the top 1 mm of water off the airwater interface throughout
all experiments. The overhead PIV cameras were made to travel downstream at the
average interfacial velocity.
For every flow condition, measurements were made of plan-view PIV, side-view
PIV, dissolved oxygen concentrations necessary to calculate the airwater gas transfer
coefficient, the wind velocity profile, fan power, air temperature, capacitance-wire
water heights at 500 Hz with a 200 m diameter insulated wire, water slope fields via
a refraction-slope method, water temperature, bulk water velocity, and time-averaged
water surface velocity via float tracers. All measurements were made at downstream
locations between 8 and 10 m. A detailed explanation of each measurement method
is given in the forthcoming paragraphs. The controlled parameters were water
height, water temperature, air temperature, fan power, bulk water velocity and timeaveraged surface velocity. Measurements of the various dependent parameters such
as PIV, dissolved oxygen concentrations, and wind velocity profiles were not made
simultaneously, but the controlled parameters were measured carefully to ensure

597

Airwater gas transfer and near-surface motions


Cameras

Airwater
interface

(b)
Depth below airwater interface

(a)

0
0.7

(c)
1.4

Ruler
45
Airwater
interface

2.1
2.8
0

25

50

75

100 125 150 175 200 225 250

Digital cameras pixel value (0255)

F IGURE 2. Schematic of the 3D-IPIV technique showing the stereoscopic cameras and the
blue stimulation light. (a) The two high-speed cameras set up in stereo configuration for threedimensional measurements. (b) Verification measurements of the visibility of flow tracers
under the water surface to our PIV cameras. (c) The method for fixing microsphere flow
tracers at known depths to make the verification measurements of tracer visibility.

reproducibility of the flow conditions. Table 4 shows the seventeen open-channel


flow conditions investigated without wind forcing. Five water depths were investigated
ranging between 5 and 13 cm, and for each of these, four water velocities were
investigated ranging from 0.017 to 0.55 m s1 . Table 5 shows the wind-driven flow
conditions, which included thirteen wind velocities and two water depths. Calculation
of the airwater friction velocity and estimation of the wind velocity at ten metres
height U10 is explained in the following paragraphs.
To characterize near-surface motions, a new particle image velocimetry (PIV)
technique was developed to allow accurate measurement of water velocity within one
millimetre of the airwater interface, termed interfacial particle imaging velocimetry
(IPIV). In addition to this technique, if the two high-speed cameras were placed in
a stereoscopic arrangement then measurement of the interface topography was made
simultaneous with the IPIV. The combination of these two techniques is termed threedimensional IPIV (3D-IPIV). Figure 2(a) shows a schematic of the 3D-IPIV version
of these techniques, wherein the entire flow channel water is made opaque with 1
part in 10 000 (w/w) water-soluble acid dye (equal parts FD&C Red 40, FD&C Blue
1, FD&C Yellow 5) and seeded with fluorescent-red microsphere flow tracers that
are stimulated by a blue overhead light. The fluorescent microspheres were measured
with a Malvern particle size analyser to have mean diameter of 42 m diameter
with standard deviation of 10 m. The cameras were mounted onto the blue light
in a stereoscopic geometry for 3D-IPIV measurements whereas they were oriented
parallel to each other for IPIV measurements. Flow tracers are only visible to the
PIV camera when they are within 1 mm of the airwater interface, even for wavy
interfaces. To ensure that flow tracers were only visible when within 1 mm of the
airwater interface, individual fluorescent microspheres were glued onto a ruler at
half-millimetre increments and the ruler was submerged in the opaque water at a
45 angle as shown in figure 2(c). Figure 2(b) shows the visibility of flow tracers

598

D. E. Turney and S. Banerjee

Id

Water
depth
h (m)

Bulk
velocity
u (m s1 )

Macroscopic
Reynolds
number, Re,
uh/

Channel width
over water
depth, wc /h

Fetch to
measurement-location
over water depth, d/h

R1
R2
R3
R4
R5
R6
R7
R8
R9
R10
R11
R12
R13
R14
R15
R16
R17

0.050
0.050
0.050
0.050
0.050
0.055
0.070
0.070
0.070
0.090
0.090
0.090
0.100
0.100
0.130
0.130
0.130

0.017
0.025
0.092
0.230
0.550
0.511
0.065
0.160
0.399
0.051
0.123
0.308
0.092
0.055
0.034
0.084
0.211

850
1250
4510
11 500
27 500
28 105
4550
11 200
27 930
4590
11 070
27 720
9200
5500
4420
10 920
27 443

14.0
14.0
14.0
14.0
14.0
12.7
10.0
10.0
10.0
7.7
7.7
7.7
7.0
7.0
5.4
5.4
5.4

160
160
160
160
160
145
114
114
114
88
88
88
80
80
62
62
62

TABLE 4. Conditions for the open-channel flow experiments. The average water velocity
was measured by a rotameter. Uncertainties for these data are always less than 5%.

Id
W1
W2
W3
W4
W5
W6
W7
W8
W9
W10
W11
W12
W13

Water depth
h (m)

Air-side friction
velocity ua (m s1 )

Air velocity
U10 (m s1 )

0.05
0.05
0.05
0.05
0.05
0.10
0.10
0.10
0.10
0.10
0.10
0.10
0.10

0.03
0.03
0.04
0.07
0.09
0.10
0.13
0.16
0.20
0.24
0.24
0.28
0.32

1.1
1.1
1.5
2.4
3.0
3.1
4.1
5.0
6.0
7.1
7.0
8.0
8.9

TABLE 5. Conditions for the wind-wave flow experiments. The fetch to measurement
location was 810 m for IPIV and 3D-IPIV, and 10 m for all other measurements. The
measurement of air-side friction velocity ua and its extrapolation to 10 m height U10 is
described in the text surrounding equation (2.4).

to the PIV cameras to have an attenuation length of 0.8 mm. The effect of the
dye on the airwater interfacial surface tension was measured to be negligible by
dynamic surface tensiometry (Turney 2009; Turney, Anderer & Banerjee 2009). The
stereoscopic calculations of water surface morphology proceeded according to Willert
& Gharib (1991), wherein each image from camera A was broken into subwindows,

Airwater gas transfer and near-surface motions

599

each of which was cross-correlated with a simultaneous image from camera B similar
to our PIV calculations. These cameras and PIV correlation methods are described
in the next paragraph. The error in the water height morphology was estimated
by comparing the r.m.s. and mean water height as measured by 3D-IPIV with that
measured by our wire-capacitance gauge, giving errors less than 10 % (Turney et al.
2009).
Two identical Kodak Motion Corder high-speed cameras were used for the side-view
PIV, overhead IPIV, and overhead 3D-IPIV measurements. Each camera had 512 480
pixel resolution, and could capture images at up to 250 frames per second. The frame
rate for each measurement was chosen to eliminate blurring of flow tracers. The
field of view for each camera was 7.9 cm 7.4 cm for all of our plan-view data
(i.e. IPIV and 3D-IPIV) and 6.0 cm 5.6 cm for most of our side-view PIV data.
The cameras were mounted parallel to each other for IPIV and also for the side-view
PIV, such that their images were digitally stitched together to make a wider view,
13.5 cm 7.4 cm and 11.8 cm 5.6 cm respectively. The stereoscopic data (3D-IPIV)
retained a field of view of 7.9 cm 7.4 cm because the two cameras fields of view
stereoscopically overlapped one another. The 3D-IPIV technique was applied only for
the wind-forced conditions, i.e. table 5, whereas the IPIV technique was only used for
the open-channel flow conditions, i.e. table 4. The side-view PIV was collected for all
flow conditions. The field of view of side view PIV was centred either on the top or
the bottom 5 cm of the water column. The purpose of viewing the bottom region was
to measure the channel-bottom friction velocity ucb . Additional side-view PIV was
collected for our higher Reynolds number flows R9, R12, and R17 with a narrower
field of view of 3.5 cm 3.2 cm per camera and narrower yet, 1.8 cm 1.9 cm, for R5
and R6. A hierarchical PIV algorithm (Willert 1997) was used for all PIV processing,
with an initial window size of 20 20 for the plan-view IPIV methods, 30 30 pixels
for the side-view methods, and 40 40 for the stereoscopic water height calculations.
A final subwindow size of 16 16 was used in all PIV methods. The 16 16
pixel subwindows were overlapped by 50 % in side-view and 3D-IPIV and 75 % in
IPIV, meaning that the final cross-correlation subwindow had a 1.42.2 mm length
scale and the velocity vectors were spaced 0.50.7 mm apart in the side-view and
3D-IPIV data and 0.7 mm in the IPIV. A sub-pixel PIV interpolation method refined
the cross-correlation displacement to one fifth of a pixel. A representative image from
the side-view dataset is shown in figure 3.
Measurement of the error in our PIV measurements was made by: (i) controlled
translation of objects at constant speed past the PIV field of view; (ii) comparing the
depth-averaged water velocity as measured by PIV with the bulk flow rate recorded by
a rotameter; (iii) comparing the IPIV and 3D-IPIV time-averaged downstream velocity
with a stop-watch timing of the downstream drift of 75 m diameter interface tracers;
and (iv) manual tracking of individual particles in the raw camera images by eye and
comparing with PIV calculations (Turney 2009; Turney et al. 2009). Each of these
PIV validation tests showed the difference between the mean velocity derived from
PIV and non-PIV methods, i.e. PIV error, to be less than 10 %. Consideration will
now be given to error in calculations made using the PIV data as input. One such
bias arises in the surface divergence calculation of (1.12) due to random error between
neighbouring PIV vectors. This error is only significant when the surface divergence
is near zero. The magnitude of this bias is calculated in the supplementary material of

600

D. E. Turney and S. Banerjee


0
Mean flow

Depth, y (cm)

1
2
3
4
Channel bottom at 5 cm depth
5

Streamwise, x (cm)

F IGURE 3. Example side-view PIV velocity field for flow condition R6. The data from our
two side-by-side high-speed cameras were digitally stitched together to create this wide field
of view.

Turney et al. (2005), and has magnitude


0
noise
=

!1/2
h(unoise (n 1, m, t) unoise (n + 1, m, t))2 /(41x2 )i
,
+ h(vnoise (n, m 1, t) vnoise (n, m + 1, t))2 /(41y2 )i

(2.1)

where unoise and vnoise are the random error in PIV velocities, n and m denote the
rows and columns of the PIV nodes, 1x and 1y are distances between PIV nodes,
and angle brackets denote an average over many samplings. Under our flow conditions
and PIV setup the magnitude of this positive bias in surface divergence ranges from
0.05 to 0.2 s1 and therefore only affects our flow conditions with the lowest 0
values. This bias will be discussed again when plots of 0 are presented. Another
particular concern for PIV error arises if the flow has significant velocity gradients
occurring on length scales smaller than the final PIV subwindow size or smaller than
the final spacing of PIV vectors, either of which will cause the PIV data to be
spatially smoothed as compared with the true flow. The sharpest gradient in velocity
is the dissipation length scale , which can be estimated using Kolmogorov theory
in our open-channel flows. PIV smoothing error for u0 and (u/x)0 was studied
in homogeneous isotropic turbulence (Lavoie et al. 2007; Adrian & Westerweel
2011), the outcome being that PIV smoothing error reduces u0 by only 5 % when
the subwindow size is ten times larger than the Kolmogorov scale. Our results in
table 6 show the Kolmogorov dissipation length scales of our open-channel flows
1/4
to range from 0.02 to 0.2 cm, as calculated by = ( 3 /) , where = u0 3 /L is
the dissipation estimated at a depth L below the airwater interface. We did not
calculate at the airwater interface because turbulence at this location is highly
inhomogeneous, and thus u03 /L is a poor estimate of there. A comparison of our
0.020.2 cm to our PIV subwindow size of 0.2 cm shows that our error in
u0 from PIV smoothing is never greater than 5 %. PIV smoothing error for u is
negligible because calculation of u is a spatial and temporal averaging of u while PIV
smoothing is similarly a spatial averaging. Smoothing error is more significant for our
calculations of 0 because they contain (u/x)0 terms that are sensitive to smoothing:
for example, Lavoie et al. (2007) found that PIV smoothing of (u/x)0 results in a

Airwater gas transfer and near-surface motions

601

15 % reduction compared to true values when the PIV subwindow size is five times
larger than the Kolmogorov dissipation length () and 30 % reduction when it is ten
times larger. This systematic lowering of 0 is significant only for our most turbulent
or windy conditions, where our 0 could be up to 15 % lower than true values. This
bias is taken into consideration with our error bars in plots of 0 in figures 6 and
14. We will return to this topic when presenting our open-channel 0 results in 3.
Turning attention now to our wind-driven flow experiments, the PIV smoothing at
wind speeds below U10 3 m s1 (i.e. below our W6 condition) is negligible because
the interfacial motions are well known to have length scales 100w /uw (Rashidi &
Banerjee 1990a,b), where w is water viscosity and uw is the friction velocity on the
water side of the airwater interface, which gives length scales much larger than our
PIV vector spacing. At our wind speeds W7 to W13 the smoothing error is determined
by viewing probability density functions for the statistic normalized by u0 , which
allows detection of smoothing error over the spectrum of length scales. PIV smoothing
is expected to bias wind conditions W7W13 low by 1050 %, with a greater effect
at greater wind speeds. A final source of error occurred in flow conditions W12
and W13, due to the collection and aggregation of flow tracers and trace surfactant
at the intense convergence zones at the spilling region of the microscale breaking
waves, which was orders of magnitude reduced compared with previous publications
mentioned in 1 but still caused 0 to be biased low. Based on our observations of
calculations of raw flow tracer images, we estimate this bias in 0 to be another 30 %
for flow conditions W12 and W13 only. A bolster of confidence in our PIV data is
that, for all our wind conditions, our mean downstream velocity measured by 3D-IPIV
agreed to within 15 % of hand-timing of 75 m diameter interface tracers as they
traveled 8.0 m downstream.
The stereoscopic calculations for the 3D-IPIV method proceeded according to
Willert & Gharib (1991) with a calibration against surfaces with known threedimensional profiles. The accuracy of the PIV was further validated by checking
agreement between vertically directed Reynolds stress in the side-view PIV data
calculated by uv at the closest node to the channel bottom as compared with ucb as
calculated from the law of the wall. These two calculations lie within 8 % of each
other over all the flows.
In total, for each flow condition, between four and eight thousand images were
captured. The side-view PIV and plan-view IPIV calculations were processed for
several thousand of these images by custom-written PIV software. PIV accuracy
was verified by moving objects past the camera at known speeds by stepper motor
control. Each PIV or IPIV image produced a 116 124 two-dimensional matrix of
velocity data. Each 3D-IPIV image comprised a 116 124 two-dimensional matrix of
velocity data along with a corresponding 116 124 two-dimensional matrix of water
surface height data. Each 3D-IPIV image required two hours of computer processing
time, and consequently, only 190 images were processed for each flow condition.
Turbulence statistics were calculated from these collections of 190 images from each
flow condition. These 190 images were well separated from each other by random
amounts of time, to eliminate temporal correlations with each other. The sample size
for each turbulence statistic was ensured large enough to reduce the variation, with
respect to sample size, to less than 5 %.
For the open-channel flow conditions the turbulent r.m.s. velocities and turbulent
integral length scale L in the bulk water at various depths below the interface were
measured by side-view PIV. The main goal was to measure L at a depth of L in
order to create the input data for the SDM model. Two hundred PIV fields like the

602

D. E. Turney and S. Banerjee

one shown in figure 3 were processed per flow condition. Convergence of u0 , v 0 and
L statistics show standard deviation of 4, 2 and 9 % respectively. Turbulent length
scales L were calculated as the average of Lvv (y) and Lvv (x) in the bulk water flow
according to
!
Z h
Z FOV
1
v(a)v(a + y)
v(a)v(a + x)
dy +
dx
(2.2)
L=
0
0
0
0
2
0 v (a)v (a + y)
FOV v (a)v (a + x)
where Lvv (y) is the left-hand term, Lvv (x) is the right-hand term, a is an arbitrary
location in the bulk water, h is the side-view PIV vertical field of view which in some
cases was the airwater interface, FOV is the side-view PIV data horizontal field of
view, y is the vertical direction and x is the downstream direction. Two hundred PIV
velocity fields were used for calculation of Lvv (y) and Lvv (x) for each flow condition.
The magnitude of Lvv (y) was not significantly different from Lvv (x), so we chose not
to multiply Lvv (x) by 2 along the lines of the theory of Tennekes & Lumley (1972,
p. 273). The turbulence intensities u0 and v 0 were calculated throughout the water
column at each PIV node in the vertical direction by use of this same side-view PIV
dataset. The friction velocity on the channel bottom ucb was calculated by fitting a
smooth boundary law of the wall,
ucb  yucb 
u =
+ 5.3,
(2.3)
ln
0.41

where u is the mean downstream velocity, v is viscosity, and y is distance from the
boundary. The data fitted the law of the wall very well including a linear viscous
layer at yucb / 1. The accuracy of the friction velocity calculations was checked
by comparing them to Reynolds stress calculations uv near the bottom of the channel.
The average difference between these two methods across all flow conditions was 8 %.
Surface divergence was calculated with the plan-view IPIV and 3D-IPIV by use of
(1.12) except in conditions with a wavy airwater interface, in which case corrections
were made for the slope of the interface.
For all our flow conditions the airwater gas transfer coefficient was calculated by
the method of McCready & Hanratty (1985), wherein the downstream gradient in
dissolved oxygen content is measured and a mass balance equation at steady state is
used to calculate the interfacial flux. The dissolved oxygen measurements for these
calculations were made at two downstream distances separated 1.33 m apart. At both
of these downstream distances water samples were continuously drawn and mixed
from three channel-bottom flush-mounted holes separated from each other spanwise
by 18 cm. A fluorescence-decay dissolved oxygen sensor measured the difference of
dissolved oxygen between the upstream and downstream distances. The measurement
system had a response time near 5 s and accuracy to resolution to 0.001 mg l1 .
A lock-in amplifier method was used wherein the sensor received water from the
downstream holes for five minutes then a valve was quickly switched to direct
water from the upstream holes to the sensor. Switching the source water location
back and forth ten times and calculating the mean difference between the upstream
and downstream water comprised the lock-in amplifier method, which achieved a
lower detection limit for the difference of dissolved oxygen concentration between the
upstream and downstream locations of 103 mg l1 . An advantage of this McCready
& Hanratty (1985) approach to measuring k as compared to a whole tank method
is its elimination of channel entrance and exit effects on the measurement, e.g. wave
splashing at the end of the wind-wave tank, Gortler eddies at the upstream flume

Airwater gas transfer and near-surface motions

603

entrance and wind-fetch dependence are all eliminated. The method isolates the
measurement of k to the downstream distance of 89 m. Verification of our method
was made by comparison of the measured k values to those from previous literature
that specifically mention continuous cleaning of the airwater interface. As seen in
figure 13, the agreement is very good, particularly considering the scatter typically
The standard deviation of k is 10 %. Derivation of the k
seen in comparisons of k.
calculation and documentation that the vertical profile of dissolved oxygen was selfsimilar at each sampling location, which allows measurements from bottom-mounted
probes to accurately represent the depth-averaged concentration, are both given in
Turney (2009).
The air velocity was measured with a Pitot tube of 3 mm diameter. Air velocity
measurements were made at 500 Hz for 30 s, at locations every 2.1 mm above the
airwater interface. For wind conditions W1 and W2 the airwater interface had
negligible roughness, which allowed the data to be fitted with the smooth log-law (2.3)
with ucb replaced by the air-side friction velocity of the airwater interface ua . For
wind conditions W3W13 the interface developed roughness, and we fit the data with
 
ua
y
+ B,
(2.4)
ln
u =
0.41
lr
where lr is the length scale for interfacial roughness and B is an empirical constant
with values taken from Schlichting (1955) for the corresponding lr . The data fit these
equations very well using the r.m.s. wave height as measured by 3D-IPIV for the value
of lr . At wind speeds W7 and above a neutral-wind Charnock parametrization (Smith
1988) with lr = 0.012u2a /g + 0.11/ua also fitted the data well, but below W7 the
Schlichting parametrization fit was markedly better. The probable reason for the good
fit of the Schlichting approach is the small wave age of our wind
waves. Our wave
ages reported in table 8 were calculated as cp /ua , where cp = g/(2) is the phase
speed of the dominant wavelength . The dominant wavelength was measured with
our refraction-slope methods described in the next paragraph. Our wave ages ranged
from 2.9 at wind condition W6 to 1.5 at W13. Open ocean wave ages are typically
an order of magnitude larger. Under our conditions the wind apparently interacts with
the roughness on the airwater interface as if it were stationary roughness. Plots of the
wind frictionvelocity fit, values of mean wavelength and more details are available in
Turney (2009).
For our wind-driven flows, images of the water surface slope were recorded with
a light-refraction technique wherein a 10 megabit digital camera viewed downward at
the airwater interface with a field of view of 14 cm 20 cm while a LCD computer
monitor was faced upward toward the digital camera from under a window on the
bottom of the water channel. The LCD monitor was made to display a gradient
in blue light intensity in the spanwise direction and a gradient in red light in the
streamwise direction. The water slope was calculated via Snells law and the water
surface morphology was calculated by integration away from the location of a wirecapacitance gauge. The technique was validated by placing lenses of known curvature
at the surface of the water or by comparison with wire-capacitance r.m.s. water height.
The dominant wavelengths of the wind waves were recorded using this refraction
data by manually measuring peak-to-peak distances in 25 digital reconstructions of the
interface morphology per flow condition. The phase speed cp of the mean wavelength
was then calculated using the equation given above. The presence and dimensions of
the capillary waves were clearly visualized with this data, as shown in the streamwise
slope data of figure 4.

604

D. E. Turney and S. Banerjee

(a)

(b)
1.0
0.8

0.4
0.2
0

Wind direction

5 cm

(c)

(d)

0.2
0.4

Streamwise slpoe

0.6

0.6
0.8
1.0

F IGURE 4. Calculations for the streamwise slope, made by the refraction-slope methods
discussed. Flow conditions and friction velocities: (a) W6, 3.1 m s1 ; (b) W8, 5.0 m s1 ;
(c) W10, 7.0 m s1 ; (d) W12, 8.9 m s1 .
0
0.2
0.4
0.6
0.8
1.0

0.5

1.0

1.5

2.0

2.5

F IGURE 5. Turbulence intensity profiles from side-view PIV. Each black line corresponds to
a different flow condition, with vL0 /ucb profiles on the left and u0L /ucb on the right. Grey areas
show the scatter of data from Nezu & Rodi (1986). Propagation of our estimated PIV error in
the vL0 /ucb and u0L /ucb calculations anticipates error less than 20 %, which is validated by the
agreement seen in this figure between our data and Nezu & Rodi (1986).

3. Results for open-channel flows


The statistical results for the open-channel flow experiments are given in table 6.
The magnitudes and vertical profiles of turbulence statistics closely agree with
established results for open-channel flow (Nezu & Rodi 1986; Nezu & Nakagawa
1993), as shown in the turbulence intensity plot in figure 5. The bulk turbulence

605

Airwater gas transfer and near-surface motions


( 105)
2.5
5 cm
7 cm
9 cm
13 cm

2.0
1.5
1.0
0.5

0.5

1.0

1.5

2.0

( 105)

2.5

F IGURE 6. Comparison of the McCready et al. (1986) surface divergence model (SDM)
against direct measurements of the interfacial transfer rate for open-channel flows with no
wind. The lines extending from the data points give the data point uncertainty. The symbols
denote water depths: diamonds, 5 cm; squares, 7 cm; triangles, 9 cm; asterisks, 13 cm.

integral length scale decreased as bulk flow velocity increased or as water depth
decreased. Turbulence intensity, u0 or v 0 , increased as bulk flow velocity increased or
as water depth decreased. The channel bottom friction velocity also increased as bulk
water velocity increased or as water depth decreased. The Reynolds number based on
channel bottom friction velocity of all water depths collapsed to a single line if plotted
versus bulk Reynolds number. The r.m.s. surface divergence at the airwater interface
and the airwater gas transfer rate both increased as the bulk water velocity increased
or as water depth decreased.
Figure 6 shows the comparison, for open-channel flow, between the McCready
et al. (1986) surface divergence model (SDM) prediction of the airwater gas transfer
coefficient, as calculated from our measurements of 0 , and our direct measurements
of k by oxygen absorption. The diffusivity of oxygen in water was determined with
a temperature dependence fit to data in Weast (1981). The error bars in figure 6
were determined by propagating the standard deviation in 0 and k measurements to
the 90 % confidence interval using Students t-distribution. The standard error was
identified as 10 % for k and as 15 % for 0 . The three rightmost data points were
given additional error in the positive direction due to PIV smoothing, as discussed
in 2. The leftmost data points contain additional error in the negative direction due
to the artificial surface divergence created by noise in the PIV calculations, as in the
discussion related to equation (2.1). Reasonable agreement is seen between the SDM
model and the data, with an empirical coefficient of 0.25.
We now discuss the comparison of the Banerjee (1990) surface divergence model
In figure 7 the predictions from the SDB
(SDB) with our direct measurements of k.
model are plotted as dashed lines and empirical measurements of the gas transfer
rate are plotted as solid lines. In figure 7(a) the turbulent Reynolds numbers in the
SDB model are calculated by use of u0L while in figure 7(b) u0i is used instead. A
which is
coefficient of 0.2 is found to best fit the SDB model to our measured k,
the same coefficient as found in Banerjee & Macintyre (2004) using estimates of
the turbulent length scale and velocity scale. The solid lines and solid squares are

2.3
2.0
1.7
1.8
3.1
2.7
2.2
3.5
3.1
2.5

4.2
3.4
2.9

R1
R2
R3
R4
R5
R6
R7
R8
R9
R10
R11
R12
R13
R14
R15
R16
R17

0.67
1.29
2.82
2.73
0.48
0.95
2.3
0.33
0.69
1.84

0.19
0.44
0.99

Side-view
PIV u0L
(cm s1 )

0.43
0.85
1.75
1.67
0.30
0.64
1.27
0.22
0.46
1.00

0.12
0.30
0.67

Side-view
PIV vL0
(cm s1 )

125
213
404
399
120
211
395
97
178
351

63
133
240

Ret

0.11
0.17
0.55
1.22
2.70
2.43
0.39
0.86
1.91
0.31
0.66
1.47
0.52
0.32
0.18
0.45
1.01

ucb
(cm s1 )

0.36
0.96

1.9
0.19
0.48
1.3
0.11
0.32
0.88

0.05
0.16
0.37

Plan-view
IPIV u0i
(cm s1 )

0.54
0.74

0.70
0.40
0.51
0.56
0.33
0.46
0.48

0.26
0.36
0.37

u0i
u0L

0.29
1.5
4.1
3.9
0.21
0.70
2.3
0.06
0.29
1.1

0.04
0.17
0.35

0 (s1 )

1.6
9.6

27
1.4
5.1
19.1
0.5
2.4
9.5

0.4
1.2
3.3

Max
(s1 )

1.5
8.1

21
1.3
4.5
14
0.4
2.1
7.8

0.4
1.3
3.2

Min
(s1 )

0.06
0.04
0.02
0.02
0.08
0.05
0.03
0.11
0.06
0.03

0.18
0.09
0.05

(cm)

0.4
0.9
4.6
11.8
24.5
21.7
2.0
5.9
13.5
1.4
4.0
8.3

0.5
1.7
5.5

Gas transfer
rate, Sc of
600 106 k
(m s1 )

TABLE 6. Statistical calculations for the open-channel flow data, where ReL is u0L L/, the turbulent Reynolds number (Ret ) at a depth L, u0i
is the turbulence intensity at the airwater interface as measured by IPIV or 3D-IPIV, and the other symbols are as described previously.
The standard deviation in these statistics is less than 15 % except for 0 being biased low in flow conditions R5, R6 and R9 due to trace
surfactants and 0 being biased high in R3, R7, R10 and R15 due to PIV noise that becomes significant when the flow is very quiescent, all
as described in the text surrounding equation (2.1).

L
(cm)

Id

606
D. E. Turney and S. Banerjee

607

Airwater gas transfer and near-surface motions


(a)

(b)

( 105)
5 cm

5 cm

1.5

7 cm

7 cm

1.0

9 cm
13 cm

9 cm
13 cm

2.5
2.0

0.5
0

10

15

20

Macroscopic Re

25
30
( 103)

10

15

20

Macroscopic Re

25
30
( 103)

F IGURE 7. Test of the Banerjee (1990) surface divergence model. The solid circles are our
The solid lines are best-fit lines to the solid circles, one line for each water
measurements of k.
height. The open squares are the theoretical predictions from the SDB model. In (a) L and u0L
are used for the SDB model, whereas in (b) L and u0i are used instead. The dashed lines are
best-fit lines to the open squares, one for each water height.

The dashed lines and open squares are predictions


our direct measurements of k.
from the Banerjee model, i.e. equation (1.11) in table 1. Reasonable agreement is
seen in figure 7(a) between the SDB model and measurements, which is remarkable
given that the model is based on turbulence values measured far from the interface.
This graph represents the first direct test of the SDB model and comparison to wellcontrolled experimental data. Figure 7(b) also shows a comparison between the SDB
except that the turbulence intensity at the
model and our direct measurements of k,
0
interface ui is used in calculating the SDB predictions instead of the turbulence
intensity at a distance of one L from the interface u0L . Better agreement is seen
between the predictions and the gas transfer rate measurements when u0i is used to
calculate the SDB model. This improvement occurs perhaps because small amounts of
surfactants are likely to be present for any experimental measurements at the airwater
interface. These surfactants, which seem almost impossible to remove fully, dampen
the interfacial velocities in comparison to the velocities predicted by the theory of
Hunt & Graham (1978) used by the SDB model. While the effect of this surface
dampening is relatively small it is not negligible, and it appears to have an effect that
is not linear with macroscopic Reynolds number. It should be noted that before each
experiment we vacuumed the airwater interface with the surface skimmer mentioned
in 2 for at least 30 min and kept the skimmer running during all experiments.
All skimmed water was disposed of, not recirculated. Previous experiments in openchannel flow conditions make no mention of efforts to clean surfactants off the
interface and probably are subject to such effects (Komori et al. 1989; Tamburrino
& Gulliver 2002).
Figure 8 shows plan-view surface divergence results from the IPIV data of openchannel flow. The data image in figure 8 are a zoomed-in portion of a full image
for flow condition R6, representing 10 % of the total data for that PIV snapshot.
The statistics of are calculated from 190 of these full images. Vectors of surface
velocity are plotted over the top as black arrows. Upwellings were identified as regions
of strongly divergent flows, often called boils on a river. Upwellings were often
surrounded by regions of downwelling (convergent flow) as seen in figure 8. The
divergence regions in upwellings usually contained smaller spatial structures indicative
of a spectrum of eddy sizes arising from the bulk turbulence.

608

30

4.8
4.6

20

4.4

10

4.2

4.0
3.8

10

3.6

20

3.4
3.2
1.5

2.0

2.5

3.0

Surface divergence (s1)

Streamwise position (cm)

D. E. Turney and S. Banerjee

30

3.5

Spanwise position (cm)

F IGURE 8. A portion of the field of view from an instant in time from our IPIV of surface
divergence data. The green area represents an upwelling. The ring of red area is downwelling
water. Velocity vectors, which were used to calculate the divergence, are overlaid as black
arrows.

Id

R6
R9
R12

Water
depth
d (cm)

Macroscopic
Reynolds
number Re

Turbulent
Reynolds
number Ret

Mean
surface
age (s)

Mean parent
upwelling
size lu (cm)

Integral
length scale
L (cm)

r.m.s.
divergence
0 (s1 )

5
7
9

28 105
27 930
27 720

399
395
351

0.5
1.1
1.8

1.3
1.7
2.1

1.8
2.2
2.5

3.9
2.3
1.1

TABLE 7. Statistics of upwellings in open-channel water flows.

To gain a better understanding of the age and mechanism of renewal of water at


the interface, movies of the flow tracers were further investigated from a Lagrangian
perspective. A randomly selected flow tracer was tracked backward in time until it
disappeared into the bulk, below 1 mm depth. The length of time for the tracer to
disappear into the bulk water is a measurement of the surface age of that particle
of interfacial water. Forty such measurements of surface age were made for flow
conditions R6, R9, and R12. These three flow conditions were chosen because they
have almost identical turbulent Reynolds numbers and macroscopic Reynolds numbers,
but their water depth varies from 5 to 9 cm. The spatial extent of the upwelling that
generated the particle of interfacial water is called the parent upwelling and was
delineated by locating the velocity convergence zones that surrounded the upwelling.
The size of this parent upwelling was measured by hand for each of the surface age
measurements as the average of its two lateral dimensions. By this method the mean
parent upwelling size was calculated for flow condition R6, R9 and R12. As seen from
the resulting data in table 7, keeping the Reynolds number the same, the mean age
of interfacial water increases as the water depth increases, and the r.m.s. surface
divergence 0 decreases as the water depth increases. This effect causes the interfacial
gas transfer rate to have a mixed scaling in which the water height h and the friction
velocity at the bottom of the channel, ucb , both affect the interfacial transfer rate.
This conclusion agrees with earlier work (Rashidi, Hetsroni & Banerjee 1991; Nezu

609

Airwater gas transfer and near-surface motions


(a) 2.5

5 cm

(b) 8

2.4 cm

7
2.0

1.5

1.0
0.5 5.0 cm

7 cm

7.0 cm
6.4 cm
5.1 cm
11.2 cm
10.0 cm

9 cm
13 cm

4
3
2
1

10

15

20

Re

25
30
( 103)

2.45 cm
4.95 cm
5.0 cm 10.0 cm
4.8 cm
2.5 cm 10.1 cm
9.88 cm
0.01 0.02 0.03 0.04 0.05 0.06 0.07

F IGURE 9. Measurements of k plotted versus the (a) macroscopic Reynolds number and (b)
the channel bottom friction velocity. Our measurements are plotted as solid circles, connected
by solid lines; data from Komori et al. (1989) are plotted as open diamonds; data from Moog
& Jirka (1999) are plotted as open triangles.

& Nakagawa 1993; Moog & Jirka 1999). The mixed scaling is best visualized in
figure 9.
Researchers making field measurements often require easier input data for predicting
airwater gas transfer rates. Turning to this task, the SDB model can be reformulated
in terms of easy-to-measure input data. The magnitude of u0L is well estimated as
an average of predictions of u0i , vi0 , and w0i from Nezu & Nakagawa (1993), giving
u0L 1.73ucb . Velocity measurements in open-channel flow consistently find that ucb
is equal to 4 % of the surface velocity us , giving u0L = 0.069us . The data of Nezu &
Nakagawa (1993) show that the integral length scale L can be approximated as 0.4h,
which was verified by our own data in table 6. The coefficient C in the SDB model as
discussed surrounding figure 7 is estimated as 0.2. Using these easy-to-measure input
values, the SDB model of equation (1.11) becomes
"
r
3/4 
2/3 #1/4
D
u
u

s
sh
sh

,
(3.1)
k = 0.037
h

which is tested against our experimental measurements of k in figure 10. The ratio
u0i /u0L is again multiplied against the surface velocity to account for the damping
effect of surfactants. Good agreement is seen between this easy-to-use (3.1) and
our gas transfer rate measurements. The advantage of this calculation over existing
parametrizations (Gulliver & Halverson 1989; Komori et al. 1989; Rashidi et al. 1991;
Moog & Jirka 1999; Nakayama 2000; Chu & Jirka 2003) is its clear mechanistic
picture, clear connection to the governing equations, and ease of use.
4. Results for wind-wave flows
Table 8 gives an overview of the main results for wind-driven flows. The
corresponding values of U10 are given in table 5. The r.m.s. surface divergence 0
values rise roughly linearly from 1 s1 at U10 = 2.5 m s1 , which is the wind
speed when wind waves are first seen on the interface, to 22 s1 at our highest
wind speed of U10 = 8.9 m s1 . The maximum and minimum for a given flow
condition were identified as the 95th percentile of the probability density function

610

D. E. Turney and S. Banerjee


3.0

( 105)
5 cm

2.5
2.0
1.5

7 cm

1.0

9 cm

0.5

13 cm

10

15

20

Macroscopic Re

25

30
( 103)

F IGURE 10. Comparison of the easy-to-measure version of the SDB model predictions of
k from (3.1) plotted as dashed lines and open squares, versus our direct measurements of k
shown by solid circles and solid lines. Here the Reynolds number is based on the average
airwater surface velocity and water depth.

of . Isolated values of reached well above 100 s1 at the higher wind speeds at
spilling regions of microscale breaking waves. Hundreds of these instantaneous values
of were double-checked manually from the raw flow tracer measurements. No errors
were found in comparison of manual PIV with the automated 3D-IPIV data. The
length scale of the field on the interface was calculated with an autocorrelation
function similar to (2.2) for turbulent integral length scales, and shows a decrease in
length scale as wind speed increases. The water height measurements were calculated
with the 3D-IPIV data, and agree with wave-gauge data to within 10 % error (Turney
2009). The last row of table 8 shows our measurements of the percentage surface
area covered by downstream velocity greater than 0.9cp , where cp is the mean wave
celerity. Wave breaking occurs when the surface velocity exceeds the wave celerity.
Therefore the data in the last row indicate the percentage area of the airwater
interface that is probably undergoing wave breaking. This measure of wave breaking
significantly increases at U10 > 4.5 m s1 , which was confirmed through refractionslope photographs and playback of the flow tracer movies (Turney 2009). At lower U10
wind speeds, between 2.5 and 4.1 m s1 , playback of the flow tracer movies did not
exhibit similar breaking phenomena or turbulent motions. Microscale waves appear on
the interface at U10 2.5 m s1 as evidenced by the r.m.s. water height but they do
not begin to break vigorously until 4.1 m s1 is reached. A change in the nature of
the wave motions appears to occur at U10 4.1 m s1 .
The divergence values reported here are consistently higher than those from PIV
studies of Turney et al. (2005) by a factor of two and also consistently higher than Xu
& Khoo (2006) by a factor of ten, but agree with the value from high-resolution sideview PIV of Peirson & Banner (2003) at their wind friction velocity of 0.35 cm s1 .
At the spilling toe of microscale breaking waves Peirson & Banner saw a localized
of 175 s1 , comparable to our measurements of in these same locations of up
to 150 s1 , which were at 99 % of the probability density function of our dataset.
The disagreement of surface divergence values with Turney et al. (2005) is due to
the surface-bound flow tracers used for their study, as discussed earlier. Surface-bound
tracers jam into each other at convergence zones and under-report the value of r.m.s.
surface divergence. The lower 0 values of Xu & Khoo (2006) perhaps came from

W1
W2
W3
W4
W5
W6
W7
W8
W9
W10
W11
W12
W13

Id

3
3
4
7
9
10
13
16
20
24
24
28
32

ua
(cm s1 )

0.11
0.13
0.26
0.71
0.95
1.7
5.0
8.9
13
19
18
20
22

3D-IPIV
0 (s1 )

0.4
0.5
1.5
3
5
7
14
27
38
55
57
62
64

Min
(s1 )

1.0
1.5
1.2
0.7
0.8
0.6
0.4

Length
scale of
(cm)

1.5
2.3
3.1
3.9
4.7
5.5

XKC06
0 (s1 )

1.4
1.6
2.5
3.8
5.3
6.7
12
20
28
38
37
49
60

Airwater
transfer Sc
of 600 k
(106 m s1 )

0.08
0.10
0.11
0.27
0.33
0.49
0.66
0.56
0.91
1.06

Max
wave
height
(cm)

TABLE 8. Statistics of our measurements in wind-driven flow conditions

0.4
0.4
1.3
3
5
7
14
25
33
48
46
49
52

Max
(s1 )

0.09
0.13
0.13
0.23
0.32
0.40
0.47
0.50
0.58
0.80

Min
trough
height
(cm)

0.02
0.03
0.03
0.06
0.10
0.13
0.18
0.16
0.24
0.32

r.m.s.
water
height
(cm)

0.00
0.01
0.02
0.09
0.35
0.35
1.1
1.5

Percentage
of u
above
0.9cp %

Airwater gas transfer and near-surface motions


611

612

D. E. Turney and S. Banerjee


(b)

20

Surface divergence (s1)

(a)

20

(d)

50

cm

4
3
0

2
1
0
0

cm

Surface divergence (s1)

Wind

(c)
6

50

F IGURE 11. Images of 3D-IPIV surface divergence for conditions (a) U10 = 3.1 m s1 ,
(b) U10 = 4.1 m s1 , (c) U10 = 7.1 m s1 and (d) U10 = 8.9 m s1 . Note the appearance of
capillary waves in (b), (c) and (d), and short-scale turbulent structure in (c) and (d). The
colour scale for (a, b) is 20 to 20 s1 and for (c, d) 50 to 50 s1 .

the relatively small 45 cm annulus diameter wind channel. Further, Xu et al.s vertical
profiles of r.m.s. velocity (surface normal velocity) have a sharp discontinuity near
the interface, which may indicate problems with either surface-identification or PIV
measurements near the interface.
Figures 11 and 12 show example images of our 3D-IPIV data at four increasing
wind speeds. The statistics of are calculated from large datasets of such images
(as described in 2). In figure 11 the three-dimensional data is flattened to two
dimensions for visualization purposes, and in figure 12 the third dimension is made
apparent. The lowest wind conditions between U10 = 1.1 and 2.5 m s1 exhibited a
relatively flat airwater interface. At these low wind speeds streamwise streaks are
seen to dominate the interfacial flow field, similar to previous experiments reported
at low wind speeds (Rashidi & Banerjee 1990a,b). For flow conditions above W4,
i.e. U10 > 2.5 m s1 , a sudden increase in r.m.s. water height and 0 occurred, waves
appeared on the interface, and the surface divergence spatial patterns appeared as a
mixture of the streamwise streaks and wave-associated patterns. Frequent interaction
between the wave motions and the streamwise streaks are seen in the 3D-IPIV movies
of near-surface flow tracers, in which momentum and energy are transferred from
wave to streamwise streak. At wind speeds above 4.1 m s1 , i.e. flow conditions
W8 and above, the streamwise streaks decreased significantly in prevalence and
the interfacial flow field became dominated by wave breaking motions. Surface
divergence appeared on the upwind side of wave crests, convergence on the downwind

613

Airwater gas transfer and near-surface motions

60

ection

dir
Wind

40
20

0
1

20

Sp 0
an
wi 1
2
se
po
sit 3
ion
4
(cm
)

40

5
6

ion (cm

osit
wise p
Stream

Surface divergence (s1)

Vertical position
(cm)

80

60
80

F IGURE 12. Three-dimensional plot of the 3D-IPIV data for an instant in time for wind
speed U10 = 8.9 m s1 . Strong convergence is seen on the front side of the wave. The crest
velocities grow larger than the crest speed and turbulent eddies appear on the downwind side
of the wave. A movie sequence of these data is available in the supplementary material.

side, and lines of convergence and divergence of magnitude 3 0 appeared on the


downwind face due to capillary waves riding the main wind waves. A movie of
3D-IPIV data of surface divergence on a microscale breaking wave is available in
the supplementary material for this paper (available at http://dx.doi.org/10.1017/jfm.
2013.435). The presence of these capillary waves was confirmed by the 3D-IPIV
data, which showed elevations of the airwater interface that clearly corresponded
to capillary waves and were spatially correlated with the lines of surface divergence.
Our capacitance-wire gauge and refraction-slope images confirmed the presence and
configuration of these capillary waves. These capillary waves riding the downwind side
of wind waves are also easily observable by eye during a slight steady wind over
water with fetch greater than 10 m.
At wind speeds U10 between 2.5 and 4.1 m s1 , i.e. wind conditions W4 to W7,
waves appeared and surface divergence zones covered the entire back side, i.e. upwind
side, of the wave crests. At these low wind speeds the flow tracer streamlines were
laminar in that they were predictable from wave to wave and mixing with the bulk
water was minimal, i.e. groups of flow tracers could be seen downwelling below 1 mm
depth at the convergent zones and re-emerging at upwellings without change to the
flow tracer dot pattern. For example, tracer patterns were typically seen to downwell
as an upwind wave crest approached them and then upwell on the back side of the
wave, i.e. the flow had a laminar character for which tracers where not mixed or
permanently sequestered into the bulk water. These same wind speeds retained strong
streamwise streaks and did not contain parasitic capillary waves on the downwind
side of wave crests until wind speeds of 4.1 m s1 and higher. A slight amount of
surfactant, which may have survived our continuous surface cleaning method, could be
responsible for the laminar wave motions seen. This surfactant effect would surely
be present in other reported experiments unless extreme care was taken to avoid trace
amounts of contaminants. Another explanation for these non-turbulent wave motions
is a transfer of momentum from the wind waves to the streamwise streaks, which
was clearly seen to occur in the movies of near-surface flow tracers. As the wind

614

D. E. Turney and S. Banerjee


10

( 105)

8
6
4
2

0.1

0.2

0.3

0.4

F IGURE 13. A plot of our measurements of k as circles alongside data from other
publications that explicitly mention cleaning of the airwater interface: the triangles show
data from Ocampo-Torres et al. (1994), the grey solid line is the trend line of the cleaninterface data from Jahne et al. (1987), and the grey dashed line is the trend line of the
cleaned-interface data from McGillis et al. (2007).

speed was increased above U10 > 4.1 m s1 , capillary waves appeared and waves
began microscale breaking, giving the surface flow motions a different character. This
conclusion is supported by the data in the last row of table 8, which shows that at
wind conditions W7 and above (i.e. for U10 > 4.1 m s1 ), a sudden increase occurs
in the fraction of the airwater surface that has velocity exceeding 0.9cp , where cp
is the crest speed of the surface waves. For these higher wind speeds, flow tracer
streamlines revealed small turbulent eddies shedding from the breaking wave crests
and stringing themselves out behind the breaking wave crests (on the back side of
the wave). The near-surface wave motions became more turbulent than seen at wind
speeds between 2.5 and 4.1 m s1 . At these higher wind speeds, flow tracers that
submerged in the spilling crest region no longer reappeared predictably on the back
side of the crest, i.e. they mixed into the bulk. The length scale of divergence patterns
decreased as the wind speed increased above 4.1 m s1 , as seen in table 8, due to
small-scale near-surface eddies becoming prevalent. The appearance of capillary waves
and small patterns in surface divergence are also seen in figures 4, 11 and 12. A
notable observation made by watching the unprocessed raw movies of flow tracers
was that capillary waves were ineffective at sequestering flow tracers down into the
bulk water, even at the higher wind velocities. In other words, capillary waves were
observed to be ineffective at creating a new interfacial water surface.
Figure 13 shows a comparison between our measurements of the airwater gas
transfer coefficient and other published datasets that explicitly mention cleaning
of the airwater interface. For our own experiments, before any measurement, we
vacuumed the airwater interface with the surface skimmer mentioned in 2 for at
least 30 min and kept the skimmer running during all experiments. All skimmed water
was disposed of, not recirculated. Comparisons of different researchers datasets of k
usually produce scatter of 200 % or more. It is promising to see agreement in figure 13
with error between datasets going back to 1987 of only +/10 %. This provides

615

Airwater gas transfer and near-surface motions


( 105)
7
7.0

5
4
3

8.9

8.0

4.1

5.0

7.1

6.0

2
1
0

( 105)

F IGURE 14. A comparison of the McCready et al. (1986) surface divergence model (SDM)
against direct measurements of the
interfacial transfer rate for wind-sheared airwater
flows. The solid grey line is k = 0.25 D 0 . Wave breaking occurs for wind speeds above
4.5 m s1 .

confidence both in our measurement methods for k and in the reproducibility of data
taken in different laboratories that take care to clean the water surface.

Figure 14 shows the SDM model predictions, i.e. k = C D 0 , using our


Horizontal error
measurements of 0 , compared to our direct measurements of k.

bars for k are the same as for figure 6. Vertical error bars for 0 in the leftmost
five data points are calculated in the manner described around equation (2.1), with
the positive bias due to artificial surface divergence generated by noise in the PIV
vectors. The vertical bars on the rightmost seven data points are discussed in 2 and
extend higher due to PIV smoothing error and
aggregation of the surface-bound flow
tracers. Due to the square root operator in 0.25 D 0 the error in 0 is much reduced.
The data points and conservative error bounds in figure 14 suggest disagreement
between the SDM model and empirical measurements. At the inception of capillary
waves, i.e. U10 4.1 m s1 , the SDM model suddenly predicts larger k than our
As the wind speed is increased above 5.0 m s1 the
direct measurements of k.

predictions of k level off and eventually fall back down to the direct measurements
This is probably due to the sudden increase in 0 arising from the generation
of k.
of capillary waves and the decreasing prevalence of capillary waves at wind speeds
higher than 5.0 m s1 . The lack of agreement between the SDM model and our
empirical measurements is larger than the measurement error.
The commonly used SDM model equation (1.10) is based on the assumptions that:
(i) all important surface divergence motions have long lifetime (low frequency) such
that steady-state equations can be used; and (ii) downwelling water is mixed with the
bulk water before it reappears at the surface. These assumptions were not always valid
in the wind-wave motions of our experiments. For example, although the capillary
waves that arise at wind speeds at or above 4.1 m s1 show strong surface divergence,
they are shown in our analysis below to be too short-lived for the low-frequency
solution to hold true. Further, the laminar wave motions at U10 < 4.1 m s1 produced
large surface divergence values but obviously did not mix the downwelling water with
bulk water. These near-surface capillary and laminar wave motions appear ineffective
for airwater gas transfer and create the hump in SDM datapoints seen in figure 14

616

D. E. Turney and S. Banerjee


1

at 4.1 m s . At higher wind speeds, above U10 of 5.0 m s1 , turbulence produced


by microscale breaking waves covers an increasing fraction of the interfacial surface
area and appears to clear away capillary waves and laminar wave motions. These
turbulent motions cause an increase in the airwater gas transfer rate as wind speed
increases above 5.0 m s1 , but the 0 value does not increase as strongly because
the contribution to 0 from capillary waves and laminar wave motions decreases as
wind speed increases. This explains the levelling off of the surface divergence at wind
speeds above 5 m s1 in figure 14.
McCready et al. (1986) suggest that the threshold frequency t to distinguish
low-frequency motions from high-frequency (ineffective) motions for a sinusoidal
2
oscillating divergence time series is t < D/lcbl
. We revisit the definition of this
threshold now, because the definition from McCready et al. does not include the
strength of the near-surface divergence motion, which seems to clearly be needed
in such a threshold. Analytical solutions to the advectiondiffusion equation for
stagnation flows provide insight on this threshold and are given in Chan & Scriven
(1970) and Brumley & Jirka (1988), both of which contain the time scale term
e

Rt

0 0
0 (t )dt

(4.1)

which is a threshold for the effectiveness of a surface divergence motion on the


airwater gas transfer rate. For a constant divergence of near-surface motions over a
period of time this criterion in (4.1) becomes > 1. (Further analysis of these
matters is provided in Turney 2009). For investigating individual near-surface divergent
motions we propose (4.1) as the criterion to determine the impact on interfacial gas
transfer, and to distinguish low- and high-frequency motions.
To explore which near-surface motions satisfy this threshold, the 3D-IPIV data
reported here allowed Lagrangian tracking of the values for small patches of surface
water (of scale 1 mm2 ) as they moved around on the airwater interface. An example
time series of this Lagrangian data is shown in figure 15, during which time a
microscale breaking wave passes through the patch of water. While the patch is on
the front side of the wind wave it experiences fast oscillations in surface divergence
due to the parasitic capillary waves that ride the downwind face of the wave. After
the wave crest passes the Lagrangian measurement patch, i.e. at the end of the time
series when the patch is on the upwind side (back side) of the wave, a persistent
upwelling occurs. The length of time that a capillary wave causes surface divergence
to remain positive at this Lagrangian patch of surface water is the lifetime of the
upwelling for the purposes of calculation of airwater gas transfer. As seen in the
rapid oscillations on the left side of figure 15, the lifetime of the capillary-wave
near-surface motion is 0.01 s and the strength of the upwelling is 10 s1 , so
the capillary near-surface motions do not satisfy the criterion > 1. However, the
non-capillary upwelling on the back side of the wave, on the right side of figure 15,
has 20 s1 and 0.05 s, so it does satisfy > 1. This provides quantitative
evidence that the parasitic capillary waves are not contributing to airwater gas transfer
whereas upwellings on the backside of microscale breaking wind waves are quite
effective.
To further investigate the matter, randomly selected flow tracers were tracked
backwards in time until they disappeared into the bulk water, in the same manner
as described in 3 for open-channel flows. The length of time until this randomly
selected tracer disappears below 1 mm into the bulk water is a measure of the age
of that parcel of surface water. For each age measurement, we also recorded the type

617

Airwater gas transfer and near-surface motions

20

1.0

10

0.5
0

0
10

0.5

20

1.0

30

1.5

Water height (cm)

Surface divergence (s1)

30

40
50

0.04

0.08

0.12

0.16

0.20

0.24

0.28

0.32

Time (s)

F IGURE 15. A time series of surface divergence values and water height as measured at a
Lagrangian patch of surface water as a microscale breaking wave passes by. At time zero the
Lagrangian surface patch is located on the downwind (front side) of the wave and capillary
wave oscillations in surface divergence have period 0.01 s. At time 0.20 s the microscale
breaking crest has passed by downstream of the surface patch and the surface divergence no
longer oscillates due to capillary waves but rather has a persistent upwelling. The wind speed
was U10 = 6.0 m s1 .

of upwelling event that generated the parcel of surface water, with categories being
a capillary wave upwelling, a laminar wave upwelling, a turbulent wave upwelling, a
weak upwelling, or a streamwise vortex upwelling. The difference between a weak
upwelling and a strong upwelling was determined by the value of in the upwelling,
< 5 s1 for a weak upwelling and > 5 s1 for a strong upwelling. Turbulent
upwellings were identified as motions unconnected to the mean wave motion, whereas
a laminar wave upwelling was a predictable motion related to the mean wave motion,
seen most often at wind conditions between W5 and W9. For wind conditions W9,
W11 and W13, we made forty of these measurements. Table 9 presents the results,
and shows that capillary waves generate renewed surface water in less than 10 % of
cases. In other words the surface divergence of capillary waves does not often meet
the > 1 criterion or affect the interfacial transfer rate. The near-surface motions
for which > 1 occurs most often is on the back side (upwind side) of the wave
crests, due to either turbulent upwellings, weak upwellings, or a streamwise vortex
generated by the microscale breaking wave crest. Table 9 also shows that turbulence
plays a dominant role at the higher wind speeds above wind condition W9, i.e. the
wind waves at wind speeds lower than W9 bring less turbulence into play than those at
higher wind speeds. These findings are supported by our images of surface divergence
in figures 11 and 12, which show smaller and more complex spatial structure in the
surface divergence patterns at higher wind speeds due to turbulence generation. Also,
the last row of table 8 shows that the fraction of the interface that experiences wave
breaking conditions increases as the wind speed increases above W8, which gives
further support to these findings. We conclude that microscale wave breaking plays
a role in creating surface divergence motions and turbulence at wind speeds above
U10 4.1 m s1 , but at wind speeds above 5.0 m s1 turbulent motions from the
breaking waves take a predominant presence on the interface rather than capillary

618
Id

W9
W11
W13

D. E. Turney and S. Banerjee


Mean
surface age
(s)

Capillary
wave
upwelling
(%)

Laminar
wave
upwelling
(%)

Strong
turbulent
upwelling
(%)

Weak
turbulent
upwelling
(%)

Streamwise
streak (%)

0.64
0.53
0.31

8
5
5

59
46
26

13
28
51

8
5
5

13
15
13

TABLE 9. Statistics of our surface water age measurements.

waves. This agrees with results from other studies, where side-view PIV calculations
of vorticity (Siddiqui et al. 2004; Siddiqui & Loewen 2007) and plan-view infrared
images of the airwater interface (Zappa et al. 2004) both found that the area covered
by turbulent wakes increased dramatically as wind speeds rose through the range
U10 from 6 to 12 m s1 . We additionally see that at U10 wind speeds between 2.5
and 4.1 m s1 the wind waves and surface divergence motions have a laminar wave
motion character that is different from the motions seen at higher wind speeds, and
streamwise streaks have a major presence at these low wind speeds. At very low
wind speeds U10 < 2.5 m s1 the important near-surface motions are dominated by
streamwise streaks.
5. Discussion and analysis
The results above suggest that more information is needed than just 0 and D to
calculate the airwater gas transfer rate. A Lagrangian time series of , or alternatively
the lifetime of a steady divergent motion, is needed to determine if the motion
will satisfy (4.1) or the > 1 criterion respectively. This result suggests that a
combination of the SR and SD models will be more accurate than either one alone.
Previous approaches to this issue (McCready et al. 1986; Tamburrino & Gulliver
2002) used power spectra in frequency space to analyse the data. However, since
power spectra do not differentiate between the magnitude of an oscillation and its
intermittency of occurrence, the power spectrum approach does not appear to address
the important physical phenomena. A more explicit description of the near-surface
motions, including intermittency, is needed.
We now generate a new model of airwater gas transfer that is a union of the
surface renewal and the surface divergence models. The new model assumes the
important near-surface motions to be upwellings of strength 0 separated in time
according to a Poisson distribution with mean time equal to our measurements of
the Lagrangian mean age of surface water (as given in table 9). This surface
water age is equivalent to the mean surface renewal time from Danckwerts
(1951). Our calculation of a gas transfer coefficient takes the same route as (4)
and (5) in Danckwerts original paper (Danckwerts 1951), which begins with Higbies
(1935) equation of the instantaneous concentration profile in the y-direction at time
subsequent to a surface renewal event,


c cbw
y
= erf
,
(5.1)
Scba cbw
4D

Airwater gas transfer and near-surface motions

619

which with equation (1.3) gives the instantaneous flux of gas across the interface at
time as
r
D
Fi = (Scba cbw )
.
(5.2)

Danckwerts (1951) used this result along with a Poisson probability density
distribution for the values of that occur at the airwater interface
=

e/
,

(5.3)

where is the mean age of the interface since the last renewal, or in other words
the mean surface renewal time. Calculation of the rate of time- and space-averaged
interfacial transfer Fi is obtained by weighted integration of Fi over the entire the
probability density function of surface ages, which gives
r
Z
Z /
De
D

d = (Scba cbw )
.
(5.4)
Fi =
Fi d = (Scba cbw )


0
0

The result of (5.4) is the mass transfer coefficient k = D/ previously given in


(1.7). Danckwerts approach assumes that each renewal event completely renews the
near-surface region with fresh bulk water. Our model proposes that each upwelling
event begins with the gas concentration in equilibrium with a stagnation flow of
strength 0 followed by a period of time during which the flow is stagnant and the
gas concentration adjusts accordingly. This type of flow motion is supported by our
Lagrangian IPIV data and the findings of numerical simulations from Kermani et al.
(2011). The gas concentration profile along the y-direction for an upwelling of strength
0 is given by Chan & Scriven (1970) as
r !
0
c cbw
,
(5.5)
= erf y
Scba cbw
2D
which is equal to Higbies equation for gas concentration in a stagnant liquid at
time = 1/2. We model the interfacial transfer process by combining the Higbie
surface renewal equation with the surface divergence equation of (5.5), assuming all
upwellings to instantaneously bring the near-surface concentration profile equal to (5.5)
after which the concentration profile decays according to (5.1). Addition of 1/(2) to
in (5.1) is the mathematical representation of this model and gives the interfacial
flux as
v
u
D
u
,
Fi = (Scba cbw )u 
(5.6)
1
t
+
2
which is inserted into the same integration of equation (5.4) to give
/
Z
Z
De

Fi =
Fi d = (Scba cbw )
d.

(
+ 1/(2))
0
0

(5.7)

620

D. E. Turney and S. Banerjee


( 105)
8
7
6
5
4
3
2

Open-channel flow

1
0

Wind-sheared flow
1

7
8
( 105)

F IGURE 16. Comparison of (5.8), a unification of the surface divergence and renewal models
against direct measurements of k.
The solid grey
of the airwater gas transfer coefficient k,
line is the one-to-one agreement line. Black lines extending from the data points give our
measurement error.

Equation (5.7) can be solved by use of an integrand substitution n = + 1/(2) to


generate the following prediction for the airwater gas transfer coefficient
s
!
r
D
1
0
k =
.
(5.8)
e1/2 erfc

2 0
The input data required to calculate k with (5.8) are 0 values from tables 6 and 8,
and values from tables 7 and 9. Figure 16 shows predictions of k calculated from
(5.8) compared to our direct measurements of k from oxygen absorption in the flow
channel. Figure 16 tests the new theory in both wind-sheared and open-channel flows.
Reasonable agreement is found and no corrective coefficient of proportionality has
been placed in front of (5.8). The model is expected to perform well in a wide variety
of flow conditions because its basis of derivation is valid across a wide range of flows.
This model uses simple mean values for these near-surface mixing events, 0 and ,
therefore further work is needed to consider the effect of the spectrum of 0 and
values that occur, which could be the source of minor error that the model shows in
figure 16.
6. Conclusions
The relationship between the near-surface motions and the airwater gas transfer
coefficient has been analysed in the light of new spatially resolved measurements
of , k and surface topography. The surface divergence theories of McCready et al.
(1986) and Banerjee (1990) agree with our measurements in open-channel flows in
the absence of wind shear. A mixed scaling of ucb with h is found to be necessary
to relate the 0 and k values to flow conditions. A new version of Banerjees surface
divergence model, phrased in terms of easily measured parameters, is shown to give
good prediction of airwater gas transfer coefficients in open-channel water flows.
Measurements of and k in wind-sheared flows with microscale wave breaking
indicate that the inception of capillary waves, laminar wind-wave motions and
turbulent wakes cause the relation between 0 and k to be more complex than the

Airwater gas transfer and near-surface motions

621

typically used SDM theory of McCready et al. The lifetime of a near-surface surfacenormal motion in a capillary wave or laminar wave is found to be too short to affect
the interfacial gas transfer rate. It appears that > 1 is the time scale criterion for
motions to be effective in airwater gas transfer, which requires information on both
surface divergence strength and upwelling persistence time, i.e. the surface renewal
time. Capillary waves arise at wind speeds U10 above 4.1 m s1 , causing an increase
in 0 but not affecting k as much as predicted by the SDM model because these
capillary motions do not meet the > 1 criterion. Oscillatory laminar wave motions
at U10 > 2.5 m s1 showed a similar behaviour, and were not as effective at mixing
near-surface water with bulk water as were turbulent motions at wind speeds above
5.0 m s1 . Wind waves at U10 between 2.5 and 4.1 m s1 were found to be a
mixture of laminar wave motions and streamwise streaks, which generate much less
turbulence and crest breaking than at higher wind speeds. It is unclear from our data if
these laminar wind-wave motions were a result of trace surfactants that survived our
continuous surface cleaning, or if they are due to frequent transfer of energy from the
wind waves to streamwise streaks. At wind speeds above 5.0 m s1 , crest breaking
and turbulent wakes markedly increase in presence on the interface, yet the r.m.s.
surface divergence levels off because the presence of capillary waves and laminar wave
motions decreases markedly due to destruction by the wave breaking and turbulent
wake motions of microscale breaking waves. The surface divergence motions at the
higher wind speeds (U10 > 5.0 m s1 ) are more efficient at causing airwater gas
transfer because their divergence meets the > 1 criterion more frequently than do
the motions at lower wind speeds. These findings agree with other recent studies that
found increase in the frequency of crest breaking and increase in the presence of
vorticity near the airwater interface for wind speeds 612 m s1 (Siddiqui et al.
2004; Zappa et al. 2004; Siddiqui & Loewen 2007). For wind-driven flows, in general,
analysis of our results suggests that the region behind the wave crest is the site of
upwellings that persist long enough to satisfy > 1 and cause significant gas transfer,
whereas the parasitic capillary waves on the frontside of the waves are ineffective,
which agrees with surface-water temperature patterns from infrared imagery of the
aforementioned studies.
These findings suggest that surface divergence strength and surface renewal
lifetime should both be included in any predictive model for the airwater gas
transfer rate. A quantitative model is presented here that combines elements of surface
renewal and surface divergence theories. Predictions from this first-principles model
are found to agree with measurements of k without requiring use of corrective
coefficients.
Acknowledgements

We gratefully acknowledge financial support from US National Science Foundation


grant CTS-055333, and from US Department of Energy grant DE-FG03-85ER13314.
We also thank I. Leifer, B. Piorek, C. Mathpati, A. Anderer and M. Zavrel for helping
with some of the experiments and in obtaining some of the data.
Supplementary movies

Supplementary movies are available at http://dx.doi.org/10.1017/jfm.2013.435.

622

D. E. Turney and S. Banerjee


REFERENCES

A DRIAN, R. & W ESTERWEEL, J. 2011 Particle Image Velocimetry. Cambridge University Press.
BANERJEE, S. 1990 Turbulence structure and transport mechanisms at interfaces. In Ninth
International Heat Transfer Conference, Keynote Lectures, vol. 1, pp. 395418. Hemisphere
Press.
BANERJEE, S., L AKEHAL, D. & F ULGOSI, M. 2004 Surface divergence models for scalar exchange
between turbulent streams. Intl J. Multiphase Flow 30 (78), 963977.
BANERJEE, S & M ACINTYRE, S. 2004 The airwater interface: turbulence and scalar exchange. In
Advances in Coastal and Ocean Engineering (ed. J. Grue, P. L. F. Liu & G. K. Pedersen),
pp. 181237. World Scientific.
BANERJEE, S., S COTT, D. S. & R HODES, E. 1968 Mass transfer to falling wavy liquid films in
turbulent flow. Ind. Engng Chem. Fundam. 7 (1), 2227.
B ELANGER, T. V. & KORZUN, E. A. 1991 Critique of floating-dome technique for estimating
reaeration rates. J. Environ. EngngASCE 117 (1), 144150.
B RUMLEY, B. H. & J IRKA, G. H. 1987 Near-surface turbulence in a grid-stirred tank. J. Fluid
Mech. 183, 235263.
C HAN, W. C. & S CRIVEN, L. E. 1970 Absorption into irrotational stagnation flow: a case study in
convective diffusion theory. Ind. Engng Chem. Fundam. 9 (1), 114120.
C HU, C. R. & J IRKA, G. H. 1992 Turbulent gas flux measurements below the airwater interface of
a grid-stirred tank. Intl J. Heat Mass Transfer 35, 19571968.
C HU, C. R. & J IRKA, G. H. 2003 Wind and stream flow induced reaeration. J. Environ.
Engng-ASCE 129 (1), 11291136.
DANCKWERTS, P. V. 1951 Significance of liquid-film coefficients in gas absorption. Ind. Engng
Chem. 43 (6), 14601467.
D E A NGELIS, V., L OMBARDI, P., A NDREUSSI, P. & BANERJEE, S. 1999 Microphysics of scalar
transfer at airwater interfaces. In Proceedings of the IMA Conference: Wind-over-Wave
Couplings, Perspectives and Prospects (ed. S. G. Sajjadi, N. H. Hunt & J. C. R. Thomas),
pp. 257271. Oxford University Press.
DVORAK, B. I., L AWLER, D. F., FAIR, J. R. & H ANDLER, N. E. 1996 Evaluation of the Onda
correlations for mass transfer with large random packings. Environ. Sci. Technol. 30 (3),
945953.
F ORTESCUE, G. & P EARSON, J. R. A. 1967 On gas absorption into a turbulent liquid. Chem.
Engng Sci. 22 (9), 1163&.
F ULGOSI, M., L AKEHAL, D., BANERJEE, S. & D E A NGELIS, V. 2003 Direct numerical simulation
of turbulence in a sheared airwater flow with a deformable interface. J. Fluid Mech. 482,
319345.
G ULLIVER, J. S. & H ALVERSON, M. J. 1989 Airwater gas transfer in open channels. Water
Resour. Res. 25 (8), 17831793.
H ERLINA, N. & J IRKA, G. H. 2008 Experiments on gas transfer at the airwater interface induced
by oscillating grid turbulence. J. Fluid Mech. 594, 183208.
H IGBIE, R. 1935 The rate of absorption of a pure gas into a still liquid during short periods of
exposure. Trans. Am. Inst. Chem. Engrs 31, 365.
H O, D. T., L AW, C. S., S MITH, M. J., S CHLOSSER, P., H ARVEY, M. & H ILL, P. 2006
Measurements of airsea gas exchange at high wind speeds in the southern ocean:
implications for global parameterizations. Geophys. Res. Lett. 33 (17), 321327.
H UNT, J. C. R., B ELCHER, S., S TRETCH, D., S AJJADI, S. & C LEGG, J. 2010 Turbulence and wave
dynamics across gasliquid interfaces. In The 6th International Symposium on Gas Transfer at
Water Surfaces. Kyoto, Japan.
H UNT, J. C. R. & G RAHAM, J. M. R. 1978 Free-stream turbulence near plane boundaries. J. Fluid
Mech. 84, 209235.

J AHNE
, B. & H AUECKER, H. 1998 Airwater gas exchange. Annu. Rev. Fluid Mech. 30, 443468.

J AHNE
, B., M UNNICH
, K. O., B OSINGER
, R., D UTZI, A., H UBER, W. & L IBNER, P. 1987 On the
parameters influencing airwater gas exchange. J. Geophys. Res. 92 (C2), 19371949.
K ERMANI, A., K HAKPOUR, H. R., S HEN, L. & I GUSA, T. 2011 Statistics of surface renewal of
passive scalars in free-surface turbulence. J. Fluid Mech. 678, 379416.

Airwater gas transfer and near-surface motions

623

KOMORI, S., NAGAOSA, R. & M URAKAMI, Y. 1993 Turbulence structure and mass transfer across a
sheared airwater interface in wind-driven turbulence. J. Fluid Mech. 249, 161183.
KOMORI, S., YASUHIRO, M. Y. & H IROMASA, U. 1989 The relationship between surface renewal
and bursting motions in an open-channel flow. J. Fluid Mech. 203, 103123.
K UMAR, S. & BANERJEE, S. 1998 Development and application of a hierarchical system for digital
particle image velocimetry to free-surface turbulence. Phys. Fluids 10 (1), 160177.
L AM, K. & BANERJEE, S. 1992 On the conditions of streak formation in bounded flows. Phys.
Fluids A4, 306320.
L AMONT, S. & S COTT, D. 1970 An eddy cell model of mass transfer into the surface of a turbulent
liquid. AIChE J. 16 (4), 513519.
L AVOIE, P, AVALLONE, G., D E G REGORIO, F., ROMANO, G. P. & A NTONIA, R. A. 2007 Spatial
resolution of PIV for the measurement of turbulence. Exp. Fluids 43 (1), 3951.
L AW, C. N. S. & K HOO, B. C. 2002 Transport across a turbulent airwater interface. AIChE J. 48
(9), 18561868.
L AW, C. N. S., K HOO, B. C. & C HU, T. C. 1999 Turbulence structure in the immediate vicinity of
the shear-free airwater interface induced by a deeply submerged jet. Exp. Fluids 27 (4),
321331.
L EWIS, W. K. & W HITMAN, W. G. 1924 Principles of gas absorption. Ind. Engng Chem. 16, 1215.
L OMBARDI, P., D E A NGELIS, V. & BANERJEE, S. 1996 Direct numerical simulation of
near-interface turbulence in coupled gasliquid flow. Phys. Fluids 8, 1643.
M AGNAUDET, J. & C ALMET, I. 2006 Turbulent mass transfer through a flat shear-free surface.
J. Fluid Mech. 553, 155185.
M C C READY, M. J. & H ANRATTY, T. J. 1985 Effect of air shear on gas-absorption by a liquid-film.
AIChE J. 31 (10), 20662074.
M C C READY, M. J., VASSILIADOU, E. & H ANRATTY, T. J. 1986 Computer-simulation of turbulent
mass-transfer at a mobile interface. AIChE J. 32 (7), 11081115.
M C G ILLIS, W. R., DACEY, J. W. H., WARE, J. D., H O, D. T., B ENT, J. T., A SHER, W. E.,
Z APPA, C. J., R AYMOND, P. A., WANNINKHOF, R. & KOMORI, S. 2007 Airwater flux
reconciliation between the atmospheric CO2 profile and mass balance techniques. In Transport
at the Air Sea Interface (ed. C. S. Garbe, R. A. Handler & B. Jahne), pp. 181192.
Springer.
M C K ENNA, S. P. & M C G ILLIS, W. R. 2004 The role of free-surface turbulence and surfactants in
airwater gas transfer. Intl J. Heat Mass Transfer 47 (3), 539553.
M OOG, D. B. & J IRKA, G. H. 1999 Stream reaeration in non-uniform flow: macroroughness
enhancement. J. Hydraul. Engng 125 (1), 1116.
NAKAYAMA, T. 2000 Turbulence and coherent structures across airwater interface and relationship
with gas transfer. PhD Dissertation, Kyoto University.
N EZU, I. & NAKAGAWA, H. 1993 Turbulence in Open-Channel Flows. Balkema.
N EZU, I. & RODI, W. 1986 Open-channel flow measurements with a laser Doppler anemometer.
J. Hydraul. Engng 112 (2), 335355.
O CAMPO -T ORRES, F. J., D ONELAN, M. A., M ERZI, N. & J IA, F. 1994 Laboratory measurements
of mass-transfer of carbon-dioxide and water-vapour for smooth and rough flow conditions.
Tellus B 46 (1), 1632.
P EIRSON, W. L. & BANNER, M. L. 2003 Aqueous surface layer flows induced by microscale
breaking wind waves. J. Fluid Mech. 479, 138.
P ICHE, S., G RANDJEAN, B. P. A. & L ARACHI, F. 2002 Reconciliation procedure for gasliquid
interfacial area and mass-transfer coefficient in randomly packed towers. Ind. Engng Chem.
Res. 41 (16), 49114920.
R ASHIDI, M. & BANERJEE, S. 1990a Streak characteristics and behaviour near wall and interface in
open channel flows. Trans. ASME J. Fluids Engng 112 (2), 164170.
R ASHIDI, M. & BANERJEE, S. 1990b The effect of boundary-conditions and shear rate on streak
formation and breakdown in turbulent channel flows. Phys. Fluids A 2 (8), 18271838.
R ASHIDI, M., H ETSRONI, G. & BANERJEE, S. 1991 Mechanisms of heat and mass-transport at
gasliquid interfaces. Intl J. Heat Mass Transfer 34 (7), 17991810.
S CHLICHTING, H. 1955 Boundary-layer Theory. McGraw-Hill.

624

D. E. Turney and S. Banerjee

S IDDIQUI, M. H. K. & L OEWEN, M. R. 2007 Characteristics of the wind drift layer and microscale
breaking waves. J. Fluid Mech. 573, 417456.
S IDDIQUI, M. H. K., L OEWEN, M. R., A SHER, W. E. & J ESSUP, A. T. 2004 Coherent structures
beneath wind waves and their influence on airwater gas transfer. J. Geophys. Res. 109 (C3).
S MITH, S. D. 1988 Coefficients for sea surface wind stress, heat flux, and wind profiles as a
function of wind speed and temperature. J. Geophys. Res. 93 (C12), 1546715472.
TAMBURRINO, A. & G ULLIVER, J. S. 2002 Free-surface turbulence and mass transfer in a channel
flow. AIChE J. 48 (10), 27322743.
TAMBURRINO, A. & G ULLIVER, J. S. 2007 Free-surface visualization of streamwise vortices in a
channel flow. Water Resour. Res. 43, 9.
T ENNEKES, J. L. & L UMLEY, N. H. 1972 A First Course in Turbulence. MIT Press.
T HEOFANOUS, T. G., H OUZE, R. N. & B RUMFIELD, L. K. 1976 Turbulent mass transfer at free,
gasliquid interfaces, with applications to open-channel, bubble and jet flows. Intl J. Heat and
Mass Transfer 19 (6), 613624.
T SUMORI, H. & S UGIHARA, Y. 2007 Lengthscales of motions that control airwater gas transfer in
grid-stirred turbulence. J. Mar. Syst. 66 (14), 618.
T URNEY, D. E. 2009 Improved understanding of airwater transfer of volatile chemicals. PhD
dissertation. Donald Bren School of Environmental Science and Management, Santa Barbara,
University of California.
T URNEY, D. E., A NDERER, A. & BANERJEE, S. 2009 A method for three-dimensional interfacial
particle image velocimetry (3D-IPIV) of an airwater interface. Meas. Sci. Technol. 20 (4).
T URNEY, D. E. & BANERJEE, S. 2008 Transport phenomena at interfaces between turbulent fluids.
AIChE J. 54, 344349.
T URNEY, D. E., S MITH, W. C. & BANERJEE, S. 2005 A measure of near-surface fluid motions that
predicts airwater gas transfer in a wide range of conditions. Geophys. Res. Lett. 32 (4).
WANNINKHOF, R. H. 1992 Relationship between wind-speed and gas-exchange over the ocean.
J. Geophys. Res. 97 (C5), 73737382.
WANNINKHOF, R., S ULLIVAN, K. F. & T OP, Z. 2004 Airsea gas transfer in the southern ocean.
J. Geophys. Res. Oceans 109 (C8).
W EAST, R. C. 1981 Handbook of Chemistry and Physics. CRC Press.
W ILLERT, C. 1997 Stereoscopic digital particle image velocimetry for application in wind tunnel
flows. Meas. Sci. Technol. 8 (10), 14651479.
W ILLERT, C. E. & G HARIB, M. 1991 Digital particle image velocimetry. Exp. Fluids 10 (4),
181193.
X U, Z. F. & K HOO, B. C. 2006 Mass transfer across the Turbulent gas-water interface. AIChE J. 52
(8), 33633374.
Z APPA, C. J., A SHER, W. E., J ESSUP, J., K LINKE, S. R. & L ONG, S. R. 2004 Microbreaking and
the enhancement of airwater transfer velocity. J. Geophys. Res. 109, C8.

You might also like