Free access
TECHNICAL PAPERS
Dec 15, 2010

Effects of Hydrodynamic Conditions on DO Transfer at a Rough Sediment Surface

This article has been corrected.
VIEW CORRECTION
Publication: Journal of Environmental Engineering
Volume 137, Issue 1

Abstract

A numerical model was developed to calculate the rate of dissolved-oxygen (DO) diffusion across a sediment surface taking into account the surface roughness and biochemical reactions of the sediment. Estimates of DO transfer rate from the model were compared with results from laboratory experiments conducted in a rectangular flume using roughness elements. In experiments, there was maximum value for the nondimensionalized DO transfer rate (Stanton number, St ) in the transitional region of surface roughness, in which the mass flux was two to five times larger than that of the smooth surface. The reproducibility of the experimental results by numerical analysis was significantly improved by including terms for flushing frequency of water in cavities between the roughness elements and for nonsteady variations in the diffusion rate due to step changes in DO concentration in the flushed region. A simple method to estimate enhancement effect for St caused by nonsteady variations was also presented.

Introduction

In aquatic ecosystems, dissolved oxygen (DO), the strongest biologically useful oxidant (Bakker and Helder 1993; Glud et al. 1994; Steinberger and Hondzo 1999), is critical to photosynthesis, respiration, and metal oxidation. Therefore, accurate estimation of the DO balance in a water column is indispensable to understand material cycles in a water body. In general, DO is supplied by reaeration at the air-water interface or photosynthesis by submerged plants and phytoplankton. Oxygen is consumed by living organisms and chemical processes and sediment oxygen demand (SOD) is a dominant sink term of DO (Nakamura and Stefan 1994; Steinberger and Hondzo 1999). SOD controls the oxidation-reduction potential in the surface layer of the sediment and material balance near the sediment-water interface. Although SODs have been studied by many researchers, the accurate estimation of SOD is still difficult (Arega and Lee 2005). Estimation is complicated by benthic animal activity that may enhance diffusion by bioturbation or increase of sediment surface area by burrowing (Revsbech and Jørgensen 1986) and surface roughness is equally important in controlling the SOD value.
After the 1980s, the hydrodynamic conditions above the sediment were shown to affect mass transfer rates at the sediment-water interface (e.g., Belanger 1981; Inoue et al. 2000; Arega and Lee 2005). This is due to the formation of a diffusive boundary layer (DBL), typically less than 1 mm thick, where solute concentrations change abruptly (Santschi et al. 1983; Revsbech and Jørgensen 1986). Such a boundary layer can control scalar transport (Steinberger and Hondzo 1999) and influence some biochemical reactions occurring just below the sediment-water interface. Surface roughness affects the formation of a boundary layer and rate of solute diffusion within it.
The effect of surface roughness on the rate of heat and mass transfer at the solid-liquid interface has been investigated (e.g., Yaglom and Kader 1974). It has been reported that the heat and mass transfer rates over a rough surface are generally two to four times larger than those over a smooth surface (Dipprey and Sabersky 1963; Dawson and Trass 1972; Zhao and Trass 1997). In the water environment field, surface roughness was found to enhance mass transfer from/to permeable sediments (e.g., Forster et al. 1996) and in coral reef communities (e.g., Bilger and Atkinson 1992; Falter et al. 2005); however, fewer enhancement effects have been reported for less permeable sediments such as clays (Røy et al. 2005).
Less permeable sediments often form in eutrophic lakes and coastal seas because sediment particles are very fine and each particle does not become an element of hydrodynamical roughness. However, surface irregularities, mounds, and nodules formed by the activities of benthic animals and the wave motion are often observed. These sediment surfaces are sometimes considered to be hydrodynamically rough and an increase in mass transfer rate is possible. However, there are few studies on the effects of roughness on material transfer rates at the sediment-water interface. Dade (1993) theoretically modeled and analyzed mean velocity and dissolved-substance concentration profiles just above a rough sediment-water interface and solute fluxes across the interface by a closure scheme; however, in this model, substance concentrations at the sediment-water interface were treated as zero. The concentration of reactive substances such as DO at the sediment-water interface could not be determined a priori because the concentration is affected both by physical conditions above the sediment-water interface and biochemical reactions in the sediment.
Solute transfer rates at the sediment-water interface are often estimated by multiplying the diffusion coefficient and concentration gradient at the interface; however, this method almost always underestimates the total transfer rate (e.g., Berg et al. 2003; Kuwae et al. 2006). Although possible explanations, such as an underestimation of sediment surface area and the effects of pore water advection induced by horizontal pressure gradients have been suggested (Røy et al. 2005), there is no accepted explanation.
The present study is an experimental and theoretical examination of the effects of surface roughness on the DO diffusion rate. The model includes hydrodynamic conditions above the sediment-water interface and the biochemical reactions in the sediment. The model proposed by Dade (1993) was used to calculate the mass flux in the DBL and the sedimentary kinetic model proposed by Nakamura and Stefan (1994) was used to evaluate the DO concentration at the sediment-water interface. Moreover, the flushing interval and the nonsteady process related to “vortex shedding” are considered to be the cause of differences between experimental results and calculation results. For this reason, a simple equation for modification of steady solutions is proposed.

Materials and Methods

Laboratory Experiment

The experiment was performed by means of a rectangular flume system in order to investigate the effects of surface roughness on diffusion of DO across the sediment-water interface. The experimental devices involved a rectangular flume (250 cm long, 12.5 cm wide, and 15 cm deep), in which flow velocity could be controlled (Fig. 1). This rectangular flume was hermetically sealed, and the balance of dissolved gases can be calculated. The sediment was placed in a cavity (100 cm long, 12.5 cm wide, and 10 cm deep) in the middle section of the flume bed (see also Inoue and Nakamura 2009).
Fig. 1. Schematic view of the rectangular flume system with recirculation
Sediment was collected with a shovel from a eutrophic riverbed and immediately transported to the laboratory. Sediment was placed in the aforementioned cavity to the same height as the flume bed. Fresh water (saturated with DO) was poured into the flume taking care not to disturb the sediment surface. Then, water was circulated at constant flow rate controlled by a pump and valve. The water temperature was maintained at 30±0.1°C by a temperature controller (ORION, LPA3, Suzaka City, Nagano, Japan). The experiment was started after confirming that benthic animals (river crabs and nereimorpha) were removed by means of tweezers as thoroughly as possible.
Seventeen experimental runs were completed to test three different (0 mm, 3 mm, and 5 mm) roughness conditions at six different flow velocities (Table 1) to test multiple roughness Reynolds numbers. Roughness was altered by inserting squares of 3 mm and 5-mm-high Perspex at intervals of 50 mm, fixed with narrow bars (0.8 mm in diameter) needled through both edges. Bandyopadhyay (1987) defined roughness as k type (the ratio of cavity length/roughness height>3 ) and d type (the ratio of cavity length/roughness height<1 ), and this experiment could be considered to be the k type. As this flume was too small for turbulence to fully develop, flow velocity distribution above the sediment section was measured by a laser Doppler velocimeter (TSI, Model9710, Shoreview, Minn.), and the relationship between the cross-sectional mean flow velocity (u¯) and the spatially averaged shear velocity (u) and equivalent sand roughness (ks) were obtained on the basis of a logarithmic profile of flow velocity (Grant et al. 1984).
Table 1. Experimental Conditions
Experiment numberRoughness(mm) CO(z=) (mmolL1) u¯ (cms1) u (cms1) ks (cm)
Ex#1None0.15–0.183.70.70.09
Ex#2None0.13–0.205.60.70.10
Ex#3None0.13–0.237.00.90.10
Ex#4None0.17–0.209.01.10.09
Ex#5None0.16–0.2110.41.30.08
Ex#630.17–0.201.20.20.40
Ex#730.21–0.233.10.70.45
Ex#830.18–0.225.01.10.55
Ex#930.18–0.196.21.00.26
Ex#1030.20–0.259.32.70.95
Ex#1130.16–0.229.43.40.34
Ex#1250.24–0.251.20.20.48
Ex#1350.22–0.243.50.70.40
Ex#1450.19–0.224.81.20.34
Ex#1550.19–0.237.11.40.31
Ex#1650.20–0.237.62.70.56
Ex#1750.17–0.2111.33.60.57
Photosynthesis was suppressed by covering the flume with a thick black curtain. DO concentration was monitored every hour during the experiment with a DO meter (DKK-TOA Corp., DO-25A, Shinjuku-ku, Tokyo) and the probe was installed at the downstream end of the flume (Fig. 1). The DO concentration and the meter calibration was checked by Winkler method at start and finish of each experimental run. After each experimental run, air bubbles were passed through the flume water to increase DO.
The overlying water in the flume was sampled from the sampling cock without being exposed to the air at approximately one-day intervals. The DO consumption rate per unit volume of the overlying water, rw , was measured using a quasi-biological oxygen demand method, at a temperature of 30°C , and the effect of respiration in the water was corrected. Using the following equation, SOD was calculated using decreasing rate of DO concentration:
V{dCO(z=)dt}=SODArwV
(1)
where V=the entire volume of recirculating water in the system; CO=DO concentration; t=time ; A=surface area of the sediment; and rw=volume specific oxygen consumption rate in the water.
The porosity and volume specific oxygen consumption rate of the sediment were measured immediately after the completion of each experiment. Known volumes of sediment layers were quarried from narrow sediment cores using end-cut syringes. Some portions of the sediment samples were dried at 60°C for 2 days and the porosity was estimated from the weight difference before and after the drying process. Other samples were used to measure the volume specific oxygen consumption rate of the sediment, rs , which was determined by the method of Hosoi et al. (1992). This method involved measuring the DO concentration of water clouded with sediment. Experimental conditions are listed in Table 1.

Theoretical Analysis

DO must traverse through two layers from the water column to the sediment, before consumption through respiration and chemical reactions in the sediment. The first layer is the DBL, located in the water immediately above the sediment surface. The second is the surface layer of the sediment, which is typically 2–3 mm deep for eutrophicated sediment. DO concentration at the sediment-water interface is affected by physical conditions above the interface and by sedimentary oxygen consumption kinetics. This was modeled by examining diffusion in each layer separately and then combining the results to calculate the diffusion rate of DO from the bulk water to the sediment.

Diffusion in the DBL

In Dade’s (1993) model, modified turbulent kinetic energy balance near the wall was used to define the eddy viscosity and the eddy diffusion coefficient. In the present study, the diffusion process in the DBL was estimated from Dade’s model with the following assumptions:
1.
Quasi-steady state is achieved, but water between roughness elements is intermittently flushed;
2.
Velocity and DO concentration are horizontally uniform in the bulk region; and
3.
The DBL is thin enough for DO consumption to be ignored.
Since the Damkohler number, rδ2/Dmax [where r=rate constant of oxygen consumption in the water; δ=boundary layer thickness; and Dmax=maximum value of DO diffusion coefficient in the DBL (Boudreau 2001)] in the experiments ( r=2.0×106s1 , δ=2.4cm , and Dmax=8.2×102cm2s1 ) was about 1.4×104(0.1) , Assumption 3 was appropriate (Boudreau 2001).
The diffusion rate, J , is described as
J=[Dzm+Dzt(z)]dC(z)dz
(2)
where Dzm=molecular diffusivity; Dzt=turbulent diffusivity; z=vertical distance from the sediment-water interface (positive upward); and C(z)=solute concentration. Since the Reynolds analogy states that turbulent Schmidt number, Sct , ( =vt/Dzt , where vt=turbulent kinematic viscosity) should be unity for the fully turbulent region of the flow, Dzt is expressed as the following equations (Dade 1993):
Dzt+={κz+(1z+δ+)22}+[{κz+(1z+δ+)22}24]1/22(atz+>10)
(3)
Dzt+=Kz+3(atz+<10)
(4)
where κ=Karman ’s constant ( =0.4 , Grant and Madsen 1986);and K=numerical constant taken to be 103 (Dade 1993). “ + ” denotes nondimensionalization with respect to the shear velocity, u , and the kinematic viscosity, v . From the integral calculus results from Eq. (2), the dimensionless form of the solute-concentration difference between the bulk region and the sediment-water interface can be expressed as
C+=1J0++A(z+)=1J+
(5)
where
C+u{C(z+)C(z+=0)}J
(6)
1J0+u{C(z+=z0+)C(z+=0)}J=0z0+1(1Sc+Dzt+ν)dz+
(7)
A(z+)u{C(z+)C(z+=z0+)}J=z0+z+1(1Sc+Dzt+ν)dz+
(8)
where Sc(=v/Dz)=Schmidt number. In this paper, z0 was assumed to be 0.1ks (Dade 1993), and v was selected to be consistent with Sc .
When a viscous sublayer thickness>roughness height, a wall will be treated as hydraulically smooth, and when the viscous sublayer<roughness height, it will be hydraulically rough. In k type roughness (cavity length/roughness height>3 , Bandyopadhyay 1987), eddies with a length scale proportional to roughness height are shed into the flow above the roughness elements. In d type roughness (cavity length/roughness height<1 ), eddy shedding from the elements into flow will be negligible (Perry et al. 1969). In the model, an appropriate treatment of diffusion mass transfer that includes the hydrodynamic effects of k type roughness is based on the cavity vortex theory of Dipprey and Sabersky (1963).
According to this theory, the water layer immediately above the sediment can be divided into two regions (see Fig. 2). One is a semistagnant layer occurring in roughness cavities (0<z<z0) , in which the mean flow velocity is 0, and the other is a bulk region just above the cavity (z>z0) , in which the mean velocity profile is approximately logarithmic. Mass transfer across the semistagnant layer is achieved from quasi-periodically flushing flow cells among roughness elements (Dade 1993). Although diffusion is the basic mass transfer process at the sediment-water interface, in rough situation, flushing of stagnant water among roughness elements is the limiting process for mass transfer. Consequently, a mass flux across the semistagnant layer per flushing, J , is expressed as
J=(Dzms)1/2{C(z+=z0+)C(z+=0)}=c1(Dzmuks)1/2{C(z+=z0+)C(z+=0)}
(9)
where s=flushing interval and ks=equivalent sand roughness. Dade (1993) did not include a value for flushing interval, s , but used the numerical constant, c1 , which is related to the flushing interval. A comparison of Eqs. (7) - (9) yields the dimensionless rough boundary flux as
J0+=u1(sScν)1/2=c1(RkSc)1/2
(10)
where Rk=roughness Reynolds number
Rk=uksν
(11)
J at the sediment-water interface is often expressed in terms of the Stanton number, St , which is a nondimensional diffusion rate relative to the mean velocity, u¯ , the mean concentration in the bulk region, C(z=) , and the concentration at the sediment-water interface, C(z=0) . From Eqs. (5) - (6), St can be formulated as
St=Ju¯{C(z=)C(z=0)}=(Cf2)1/2J+
(12)
where Cf=friction coefficient (2(u/u¯)2) . u¯ profiles were also calculated from integration of viscosity profiles using Reynolds analogy, and overall values of 2(u/u¯)2 were employed as respective Cf .
Fig. 2. Schematic representation of the cavity vortex hypothesis for diffusion between roughness elements
Black (1968) introduced useful scaling laws for the turbulent wall pressure field, and formulated the dimensionless mean frequency of vortex shedding based on experimental results
ωνu2=2πνsu2=0.056
(13)
where ω=averaged angular frequency of vortex shedding. Therefore, the application of Dade’s method in solving the DO transfer problem yields the SOD from Eqs. (5) - (12) as
St=SODu¯{Co(z=)Co(z=0)}=(Cf2)1/2{u(sScν)1/2+A(z+=)}1=(Cf2)1/2{(RkSc)1/2c1+A(z+=)}1=(Cf2)1/2{αSc1/2+A(z+=)}1
(14)
where α=numerical constant (=10.6) .

Diffusion and Biochemical Reactions in the Sediment

Bouldin (1968) analyzed diffusion of DO in the sediment on the assumption that the volume specific oxygen consumption rate of the sediment, rs , and the apparent DO diffusion coefficient in the sediment, Ds , are constant in the vertical direction. In such cases, the diffusion equation of DO in the sediment is expressed as
ϕdCOdt=Dsd2COdz2rs
(15)
where ϕ=porosity . By assuming a steady state and vertical integration of Eq. (15) in the oxic layer, the DO concentration profile in the sediment under the steady condition can be obtained, as follows:
CO(z)=12rsDsz2+{12rsDsδs+CO(z=0)δs}z+CO(z=0)
(16)
where δs=oxygen penetration depth, defined as
δs=2DsCO(z=0)rs
(17)
The validity of the assumption of steady state is discussed below. For the induction of Eq. (16), the following boundary conditions are used:
CO=CO(z=0)(atz=0)
(18)
CO=0(atz=δs)
(19)
Moreover, because SOD is the diffusion rate of DO at the sediment-water interface, SOD can be expressed as
SOD=Ds|dCO(z)dz|z=0=2DsrsCO(z=0)
(20)
rs can be experimentally evaluated as stated above.

Combination of Diffusion in the DBL and the Sediment

Eq. (14) represents the mass flux across the DBL and Eq. (20) represents the mass flux at the sediment surface. Because the two fluxes must coincide, the concentration at the interface, CO(z=0) , can be eliminated from the two equations. Finally, a formula for SOD as a function of u , CO(z=) , and the sediment parameters is obtained
2USOD3(4U2+1)SOD2+4USOD1=0
(21)
where the nondimensional parameters are defined as follows:
SOD=SOD2DsrsCO(z=)
(22)
U=u¯2CO(z=)St2Dsrs
(23)
Thus, according to this model, SOD=function of the mean velocity, u¯ ; mean DO concentration in the bulk region, CO(z=) ; shear velocity, u ; the equivalent sand roughness, ks ; and the Schmidt number, Sc . From these formulas we can estimate the DO concentration at the sediment-water interface from
CO(z=0)=DsrsSt2u¯2+CO(z=){DsrsSt2u¯2+CO(z=)}2CO(z=)2
(24)

Calculation Procedures

Inoue et al. (2000) investigated nonsteady variations in the material transfer rate at the sediment-water interface immediately after step changes in the material concentration of the overlying water and found that such variations substantially enhanced the material transfer rate until a steady condition was achieved. In this paper, this finding is adapted to mass transfers from cavities between roughness elements in which a water is flushed and renewed by vortex shedding.
We assume that an eddy renews the DBL down to a wall distance of βδd , where β=numerical constant (0<β<1) and δd=DBL thickness (Dade 1993)
δd=10νuSc0.33
(25)
The flushing interval s=simple function of u by Eq. (13)
s=112νu*2
(26)
Table 2 lists the values of δd and s for each experiment. According to Dworak and Wendt (1977), the eddy renews the boundary layer down to a wall distance between 30 and 50% of the DBL thickness, δd . In this model, we assumed that the eddy renews water down to a distance of 40% of δd (i.e., β=0.4 ).
Table 2. Values for Respective Processes Calculated from the Model
Experiment number δd (mm) s (s) c1 τd1 (s) τd2 (s)
Ex#10.171.780.270.874.38
Ex#20.181.900.270.934.65
Ex#30.141.130.310.552.77
Ex#40.110.710.330.351.75
Ex#50.100.550.330.271.35
Ex#60.6021.80.3010.654
Ex#70.171.850.590.904.53
Ex#80.110.710.830.351.75
Ex#90.120.840.550.412.06
Ex#100.050.121.680.060.30
Ex#110.040.081.140.040.19
Ex#120.5216.20.357.939.8
Ex#130.171.700.570.834.16
Ex#140.110.670.660.331.65
Ex#150.080.430.710.211.06
Ex#160.050.121.290.060.31
Ex#170.030.071.490.030.17
Using these values, nonsteady calculations were conducted immediately after the occurrence of vortex shedding. In these nonsteady calculations, steady DO concentration profiles were replaced by CO(z=) in the renewal region (z>βδd) and used as the initial DO concentration profiles. The calculation was iterated using respective profiles, which were also replaced by CO(z=) in the renewal region (z>βδd) , for t=s as the next initial condition until fluctuation of mass transfer rate achieved a quasi-steady state.
Fig. 3 is a schematic representation of the variation in the vertical DO profile in the vicinity of the sediment-water interface. Variation in the vertical DO profile before and after the occurrence of vortex shedding is explained by
1.
Before vortex shedding occurs, the DBL is nearly fully developed. The DO concentration gradient at the sediment-water interface is gentle and the rate of diffusion is small compared with Stage 2 [Fig. 3(A)].
2.
Immediately after vortex shedding, water in z>βδd has been flushed and the DO concentration recovers to CO(z=) [Fig. 3(B)].
3.
The DO concentration gradient at the interface temporarily becomes steeper, and the diffusion rate also increases for a short time [Fig. 3(C)].
4.
Subsequently, the DBL becomes thicker. The DO concentration gradient at the interface gradually reduces and the diffusion rate approaches a steady value [Fig. 3(D)].
5.
Immediately after the next vortex shedding, water in z>βδd has been flushed again and the DO concentration recovers to CO(z=) [Fig. 3(E)].
Fig. 3. Schematic representation of nonsteady variation in a vertical DO profile in the vicinity of the sediment-water interface: (A) before the occurrence of vortex shedding; (B) immediately after the occurrence of vortex shedding; (C) earlier phase (t<τd1) after the occurrence of vortex shedding; (D) later phase (t>τd1) after the occurrence of vortex shedding; and (E) immediately after the next occurrence of vortex shedding. Dotted lines show respective profiles at the previous stage.

Results

Experimental Results

Sample sediments had a porosity of approximately 0.76, and volume specific oxygen consumption rate, rs , of 69μmolcm3s1 . Sc was 301 (since water temperature was maintained at 30°C ). These parameters are used in the following analyses and discussion. In all experimental runs, most DO in the overlying water was consumed by sedimentary oxygen demand and DO concentration in the bulk region decreased with time. SOD was calculated from the rate of DO concentration decline using Eq. (1).
Fig. 4 shows the relationship between Rk and the St obtained from the experiment results, Stex . Although the calculated Stex varies widely, Stex increases with Rk under a smooth surface condition (Rk<30) ; however, Stex has a maximum value (two to five times larger than that on the smooth surface) in the buffer region between the smooth and complete rough surface (50<Rk<70) and declines in rough regions (Rk>100) . This tendency is similar to that described in previous papers (Dade 1993; Zhao and Trass 1997). In low Rk regions, kinetic viscosity—i.e., turbulent diffusivity—in the boundary layer is relatively small, and the DO concentration gradient in the DBL is relatively gentle. In addition, exchanges of water between the bulk region and the semistagnant layer are not frequent because the shear stress acting on the water between the roughness elements is relatively small. As a result, the DO transfer rate is small in regions of low Rk . As Rk gets larger, the shear stress acting on the water between the roughness elements also gets larger. Large shear stresses make the exchanges of water between the bulk region and the semistagnant layer more frequent and DO transfer increases. In the fully rough region, the effect of roughness elements becomes relatively large. Therefore, water exchanges between the bulk region and the semistagnant layer are not so frequent and water between the roughness elements become “dead water,” although kinetic viscosity and turbulent diffusivity in the boundary layer are high. Since the dead water is resistant to the diffusion of DO, the DO transfer is suppressed in the fully rough region. This tendency is qualitatively the same as that of heat (e.g., Dipprey and Sabersky 1963) and mass (e.g., Dawson and Trass 1972) transfer across the rough wall.
Fig. 4. Relationship between Rk and St for experimental results (solid circles; error bars are standard deviations) and St values for each experimental condition estimated by Dawson and Trass (1972) (triangles)
In Fig. 4, St values from Dawson and Trass (1972) are also plotted for each experimental condition, from the equation
StRStS=1.94Rk0.1Sc0.09(25<Rk<120)
(27)
where ScR=Stanton number for a rough surface and ScS=Stanton number for a smooth surface. The St values estimated by Eq. (27) are in the range of our experimental results, but do not fit experimental data. Bilger and Atkinson (1992) also proposed a model, but their values were around 0.0058, considerably larger than the Stex observed in our experiment and are not shown in Fig. 4.

Calculation Results

Fig. 5 shows an example of the fluctuation in DO diffusion at the sediment-water interface obtained from a nonsteady calculation over 10 s. In this example, experimental conditions of Ex#1 were used as initial parameters for nonsteady calculation. SOD achieves quasi-steady state after the occurrence of fourth vortex shedding and SOD is enhanced by up to almost twice the initial value, which corresponds with estimated value obtained by modeling excluding the nonsteady concept following the vortex shedding event. Time-averaged value under the quasi-steady state is 1.86 times initial value. The maxima develop approximately 0.6 s after the occurrence of each vortex shedding, and the monotonous increase of SOD in the initial stage is caused by an increase in the DO concentration gradient in 0<z<βδd , due to the rapid diffusion of DO from the renewed water in z>βδd . After that, SOD decreased monotonously, as the DO concentration gradient in 0<z<δd decreased due to the development of the DBL. The fluctuation of the DO diffusion rate is explained by variation in the vertical DO profile in the vicinity of the sediment-water interface before and after the occurrence of vortex shedding (Fig. 3).
Fig. 5. Example of SOD fluctuation immediately after the occurrence of vortex shedding obtained from nonsteady calculations. Arrows represent the occurrences of respective vortex shedding.
From these calculation results, the averaged SOD values under quasi-steady state were calculated, and the relationship between the experimental and calculation results of St are shown in Fig. 6(A). The experimental results were well correlated with calculation results (R2=0.786) , that is, the average error was about 6.5%. The importance of the respective formulated procedures stated above will be discussed below.
Fig. 6. Relationship between the experimental results and calculation results of St . Calculations were conducted as following: (A) c1 calculated by Eq. (29) and under nonsteady conditions; (B) c1=1 and under steady conditions; (C) c1 calculated by Eq. (29) and under steady conditions; and (D) c1=1 and under nonsteady conditions.

Discussion

Validity of the Assumption of a Steady State in the Sediment Model

In this paper, DO diffusion equation in the sediment, Eq. (15), was vertically integrated in the oxic layer assuming a steady state. First, the validity of the assumption is discussed. Using Eq. (17), the time scale of diffusion in the sediment, τs , is defined by
τs=δs22Ds=CO(z=0)rs
(28)
Using experimental values, τs was calculated to be 174–566 s. Over this time, DO concentration in the flume decreased 0.30.9μmolL1 , which correspond to 0.1–0.4% of the DO concentration in the bulk region. Therefore, we concluded the time scale required for DO profile to achieve a steady state in the sediment surface was short enough to assume a steady state in vertical integration procedure of Eq. (15). This implies the assumption of a steady state in the sediment model is valid.

Comparison of Experimental Results and Calculation Results for St with c1=1

Fig. 6(B) depicts the relationship between the experimental and calculation results excluding the nonsteady concept following a vortex shedding event. In these calculations, we assumed that c1=1 (as did Dade 1993). The calculation results underestimated the experimental results by approximately 64.2% [Fig. 6(B)], and the correlation coefficient was low (R2=0.501) . The low correlation may have two causes: the quantification of flushing interval, s (that is c1 ) and the effect of nonsteady concept following the vortex shedding event are discussed below.

Quantitative Formulation of the Flushing Interval of Dead Water between Roughness Elements

In this section, the effect of quantification of the flushing interval of dead water between roughness elements, s , was considered. Corino and Brodkey (1969) investigated the fluid motions of a turbulent flow near a solid boundary, and revealed that the viscous sublayer was continuously disturbed by small-scale velocity fluctuations and periodically renewed by fluid elements that penetrated into the viscous sublayer from positions away from the wall. Danckwerts (1951) introduced term “surface renewal” to describe this phenomenon. Corino and Brodkey (1969) considered these “bursts” to be the most important feature of the wall region, and the flushing frequency is believed to be critical factor governing the mass transfer at the surface of the wall. Under rough surface conditions, vortex shedding may be an important factor instead of burst phenomenon. Therefore, in this section, the effect of quantitative estimation of s and c1 on the calculation of St is discussed.
As s can be calculated by Eq. (13), DO flux, J , and St can be estimated from Eqs. (9) - (14). Fig. 6(C) shows the relation between the experimental and calculation results when Eq. (13) is used for the quantitative estimation of s , but neglecting the nonsteady concept following the vortex shedding event. Although the calculation results still underestimate experimental results by approximately 63.0%, the correlation coefficient increases, where R2=0.787 . From this result, we consider that the scatter of theoretically estimated values [shown in Fig. 6(B)] is attributable to inaccurate estimation of the flushing interval of dead water between roughness elements (i.e., the occurrence interval of the vortex shedding) through the model formulation procedure.
In this formulation, the numerical constant, c1 , can also be obtained as a simple function of Rk by Eq. (29).
c1=0.094Rk1/2
(29)
The c1 values calculated for each experimental condition are listed in Table 2.

Nonsteady Variations of DO Transfer after the Occurrence of Vortex Shedding

The averaged SOD values under the quasi-steady state were obtained from the same procedure as shown in Fig. 5. In this section, although the estimation of s enables us to quantify the c1 value, the improvement in model accuracy can be estimated by comparison of exclusion and inclusion of the adaptation of the nonsteady concept following the vortex shedding event. The relationship between experimental and calculation results for St is shown in Fig. 6(D). The underestimation of the calculation results shown in Fig. 6(B) is greatly improved, and the averaged error is about 12.6%. Overall underestimation of St can be attributed to exclusion of nonsteady variations in SOD after the occurrence of vortex shedding. However, without the quantitative formulation of s and c1 , the correlation coefficient is low (R2=0.433) .
The monotonous increase of SOD immediately after the occurrence of vortex shedding is caused by an increase in the DO concentration gradient in 0<z<βδd , due to the rapid diffusion of DO from the volume of renewed water where z>βδd . This is a temporary phenomenon and the time scale, τd1 , can be calculated from
τd1=(βδd)22(Dzm+Dzt1¯)=50β2ν2Sc0.67Dzmu2+250Kβ3νu2Sc1
(30)
where Dzt1¯=vertically averaged turbulent diffusivity in the range of 0<z<βδd . After that time, the steady decrease in SOD was caused by a decreasing DO concentration gradient in 0<z<δd due to the development of the DBL. This is also a temporary phenomenon, and the time scale, τd2 , of this phenomenon can be calculated from
τd2=δd22(Dzm+Dzt2¯)=50ν2Sc0.67Dzmu2+250Kνu2Sc1
(31)
where Dzt2¯=vertically averaged turbulent diffusivity in the range of 0<z<δd . The calculated τd1 and τd2 values for each experiment are listed in Table 2.
From the discussion, the reproducibility of the experimental results by numerical analysis is significantly improved by the quantitative formulation of the flushing frequency of the water in the cavity between the roughness elements (i.e., vortex shedding), considering nonsteady variations in the material transfer rate due to step changes in the material concentration of the overlying water. These concepts are applicable to other models describing heat and mass transfer at the solid-liquid interface.

Enhancement Coefficient for St

The numerical calculations of nonsteady variations after the occurrence of vortex shedding shown in the previous sections are complicated and impractical and a simple quantification of the enhancement effect for St is desirable. Fig. 7 shows the relationship between u and the enhancement factor, F , for St due to nonsteady variations within the range of experimental conditions in the present study (0.2cms1<u<3.6cms1) . The enhancement factor, F , is a monotonously decreasing function of u . The variation is aptly reproduced by a regression curve, which is formulated as a quadratic equation of u , as follows:
F=0.021u20.181u+1.985(0.2cms1<u<3.6cms1)
(32)
The enhancement factor, F , does not show a clear relationship with the equivalent sand roughness, ks , and other parameters. St values without nonsteady calculations can be simply revised from Eq. (32), based on bed shear stress information.
Fig. 7. Relationship between u and the enhancement factor for St , F , due to nonsteady variations

Conclusions

The effects of hydrodynamic conditions on the rate of diffusion of DO at rough sediment surfaces were tested experimentally and a model of diffusion to include surface roughness developed. In the experiment, St tended to increase with Rk in the smooth regions, had a maximum value in the transient rough region, and decreases monotonously in the fully rough region. This tendency is qualitatively the same as that of the heat and mass transfer rate across a rough wall. Flushing of the water in the cavities between the roughness elements (vortex shedding) and nonsteady variations in the diffusion rate due to step changes in the concentration of the renewed water also affect diffusion rates. Quantitative formulation of the flushing interval and the inclusion of nonsteady variations increase accuracy of the estimates of diffusion rates at the liquid-solid interface.

Notation

The following symbols are used in this paper:
A
=
surface area of the sediment [L2] ;
Cf
=
friction coefficient;
CO
=
DO concentration [ML3] ;
C(z)
=
solute concentration [ML3] ;
C(z=0)
=
concentration at the sediment-water interface [ML3] ;
C(z=)
=
mean concentration in bulk region of the water [ML3] ;
c1
=
numerical constant;
Dmax
=
maximum DO diffusion coefficient in the DBL [L2T1] ;
Ds
=
apparent DO diffusion coefficient in the sediment [L2T1] ;
Dzm
=
molecular diffusivity [L2T1] ;
Dzt(z)
=
turbulent diffusivity [L2T1] ;
Dzt1¯
=
vertically averaged turbulent diffusivity in the range of 0<z<βδd [L2T1] ;
Dzt2¯
=
vertically averaged turbulent diffusivity in the range of 0<z<δd [L2T1] ;
F
=
enhancement factor of DO transfer rate;
J
=
diffusion rate [ML2T1] ;
K
=
numerical constant;
ks
=
equivalent sand roughness [L] ;
Rk
=
roughness Reynolds number;
Rkmax
=
roughness Reynolds number at which Stth has the maximum value;
r
=
rate constant of oxygen consumption in the water [T1] ;
rs
=
volume specific oxygen consumption rate in the sediment [ML3T1] ;
rw
=
volume specific oxygen consumption rate in the water [ML3T1] ;
Sc
=
Schmidt number;
Sct
=
turbulent Schmidt number;
SOD
=
sediment oxygen demand [ML2T1] ;
SOD
=
nondimensional sediment oxygen demand;
St
=
Stanton number;
Stex
=
Stanton number obtained from the experiment results;
StR
=
Stanton number for a rough surface;
StS
=
Stanton number for a smooth surface;
Stth
=
theoretically calculated Stanton number;
s
=
flushing interval [T] ;
t
=
time [T] ;
U
=
nondimensional mean flow velocity;
u¯
=
mean velocity [LT1] ;
u
=
shear velocity [LT1] ;
V
=
entire volume of recirculating water [L3] ;
v
=
kinematic viscosity [L2T1] ;
z
=
vertical distance from the sediment-water interface [L] ;
z0
=
semistagnant layer thickness [L] ;
α
=
numerical constant;
β
=
numerical constant for the eddy renewal;
δ
=
boundary layer thickness [L] ;
δd
=
DBL thickness [L] ;
δs
=
oxygen penetration depth [L];
κ
=
Karman’s constant;
νt
=
turbulent kinematic viscosity [L2T1] ;
τd1
=
time scale of an increase in the DO concentration gradient in 0<z<βδd after the occurrence of vortex shedding [T] ;
τd2
=
time scale of the development of the DBL [T] ;
τs
=
time scale of diffusion in the sediment [T] ;
ϕ
=
porosity; and
ω
=
averaged angular frequency of vortex shedding [T1] .

Acknowledgments

We are grateful to the students of Department of Maritime Systems Engineering, Kyushu University for their technical help. The manuscript was greatly improved by valuable comments from anonymous reviewers.

References

Arega, F., and Lee, J. H. W. (2005). “Diffusional mass transfer at sediment-water interface of cylindrical sediment oxygen demand chamber.” J. Environ. Eng., 131(5), 755–766.
Bakker, J. F., and Helder, W. (1993). “Skagerrak (northeastern North Sea) oxygen microprofiles and porewater chemistry in sediments.” Mar. Geol., 111(3–4), 299–321.
Bandyopadhyay, P. R. (1987). “Rough-wall turbulent boundary-layers in the transition regime.” J. Fluid Mech., 180, 231–266.
Belanger, T. (1981). “Benthic oxygen demand in lake Apopka, Florida.” Water Res., 15(2), 267–274.
Berg, P., et al. (2003). “Oxygen uptake by aquatic sediments measured with a novel non-invasive eddy-correlation technique.” Mar. Ecol.: Prog. Ser., 261, 75–83.
Bilger, R. W., and Atkinson, M. J. (1992). “Anomalous mass transfer of phosphate on coral reef flats.” Limnol. Oceanogr., 37(2), 261–272.
Black, T. J. (1968). “An analytical study of the measured wall pressure field under supersonic turbulent boundary layers.” NASA Contractor Rep. No. CR-888, NASA, Washington, D.C.
Boudreau, B. P. (2001). “Solute transport above the sediment-water interface.” The benthic boundary layer: Transport processes and biogeochemistry, B. P. Boudreau and B. B. Jørgensen, eds., Oxford University Press, New York, 104–126.
Bouldin, D. R. (1968). “Models for describing the diffusion of oxygen and other mobile constituents across the mud-water interface.” J. Ecol., 56, 77–87.
Corino, E. R., and Brodkey, R. S. (1969). “A visual investigation of wall region in turbulent flow.” J. Fluid Mech., 37(1), 1–30.
Dade, W. B. (1993). “Near-bed turbulence and hydrodynamic control of diffusional mass transfer at the sea floor.” Limnol. Oceanogr., 38(1), 52–69.
Danckwerts, P. V. (1951). “Significance of liquid-film coefficients in gas absorption.” Ind. Eng. Chem., 43(6), 1460–1467.
Dawson, D. A., and Trass, O. (1972). “Mass transfer at rough surfaces.” Int. J. Heat Mass Transfer, 15(7), 1317–1336.
Dipprey, D. F., and Sabersky, R. H. (1963). “Heat and momentum transfer in smooth and rough tubes at various prandtl numbers.” Int. J. Heat Mass Transfer, 6(5), 329–332.
Dworak, R., and Wendt, H. (1977). “Stochastic fluctuations of mass transport through turbulent boundary layers.” Ber. Bunsenges. Phys. Chem., 81(9), 864–869.
Falter, J. L., Atkinson, M. J., and Coimbra, C. F. M. (2005). “Effects of surface roughness and oscillatory flow on the dissolution of plaster forms: Evidence for nutrient mass transfer to coral reef communities.” Limnol. Oceanogr., 50(1), 246–254.
Forster, S., Huettel, M., and Ziebis, W. (1996). “Impact of boundary layer flow velocity on oxygen utilisation in coastal sediments.” Mar. Ecol.: Prog. Ser., 143, 173–185.
Glud, R. N., Gundersen, J. K., Jørgensen, B. B., Revsbech, N. P., and Schulz, H. D. (1994). “Diffusive and total oxygen uptake of deep-sea sediments in the eastern South Atlantic Ocean: In situ and laboratory measurements.” Deep-Sea Res., 41(11–12), 1767–1788.
Grant, W. D., and Madsen, O. S. (1986). “The continental-shelf bottom boundary layer.” Annu. Rev. Fluid Mech., 18(1), 265–305.
Grant, W. D., Williams, A. J., and Glenn, S. M. (1984). “Bottom stress estimates and their prediction on the northern California continental shelf during CODE-1: The importance of wave-current interaction.” J. Phys. Oceanogr., 14(3), 506–527.
Hosoi, Y., Murakami, H., and Kozuki, Y. (1992). “Oxygen consumption by bottom sediment.” Proc., JSCE, Vol. 456/2(21), 83–92.
Inoue, T., and Nakamura, Y. (2009). “Effects of hydrodynamic conditions on sediment oxygen demand: Experimental study based on three methods.” J. Environ. Eng., 135(11), 1161–1170.
Inoue, T., Nakamura, Y., and Adachi, Y. (2000). “Non-steady variations of SOD and phosphate release rate due to changes in the quality of the overlying water.” Water Sci. Technol., 42(3–4), 265–272.
Kuwae, T., Kamio, K., Inoue, T., Miyoshi, E., and Uchiyama, Y. (2006). “Oxygen exchange flux between sediment and water in an intertidal sandflat, measured in situ by the eddy-correlation method.” Mar. Ecol.: Prog. Ser., 307, 59–68.
Nakamura, Y., and Stefan, H. G. (1994). “Effect of flow velocity on sediment oxygen demand: Theory.” J. Environ. Eng., 120(5), 996–1016.
Perry, A. E., Schofield, W. H., and Joubert, P. N. (1969). “Rough wall turbulent boundary layers.” J. Fluid Mech., 37(2), 383–413.
Revsbech, N. P., and Jørgensen, B. B. (1986). “Microelectrodes: Their use in microbial ecology.” Advances in microbial ecology, K. C. Marshall, ed., Plenum, New York, 293–352.
Røy, H., Huettel, M., and Jørgensen, B. B. (2005). “The influence of topography on the functional exchange surface of soft sediments, assessed from sediment topography measured in situ.” Limnol. Oceanogr., 50(1), 106–112.
Santschi, P. H., Bower, P., Nyffeler, U. P., Azevedo, A., and Broecker, W. S. (1983). “Estimates of the resistance to the chemical transport posed by the deep-sea boundary layer.” Limnol. Oceanogr., 28(5), 899–912.
Steinberger, N., and Hondzo, M. (1999). “Diffusional mass transfer at sediment-water interface.” J. Environ. Eng., 125(2), 192–200.
Yaglom, A. M., and Kader, B. A. (1974). “Heat and mass transfer between a rough wall and turbulent fluid flow at high Reynolds and Peclet numbers.” J. Fluid Mech., 62(3), 601–623.
Zhao, W., and Trass, O. (1997). “Electrochemical mass transfer measurements in rough surface pipe flow: Geometrically similar V-shaped grooves.” Int. J. Heat Mass Transfer, 40(12), 2785–2797.

Information & Authors

Information

Published In

Go to Journal of Environmental Engineering
Journal of Environmental Engineering
Volume 137Issue 1January 2011
Pages: 28 - 37

History

Received: Dec 8, 2009
Accepted: Jun 21, 2010
Published online: Dec 15, 2010
Published in print: Jan 2011

Permissions

Request permissions for this article.

Authors

Affiliations

Tetsunori Inoue [email protected]
Senior Researcher, Dept. of Marine Environment and Engineering, Port and Airport Research Institute, 3-1-1, Nagase, Yokosuka 239-0826, Japan (corresponding author). E-mail: [email protected]
Yoshiyuki Nakamura [email protected]
Distinguished Researcher, Port and Airport Research Institute, 3-1-1, Nagase, Yokosuka 239-0826, Japan. E-mail: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share