Free access
OTHER TECHNICAL PAPERS
Oct 15, 2009

Effects of Hydrodynamic Conditions on Sediment Oxygen Demand: Experimental Study Based on Three Methods

Publication: Journal of Environmental Engineering
Volume 135, Issue 11

Abstract

The quantitative effects of hydrodynamic conditions on sediment oxygen demand (SOD) under smooth surface conditions were investigated using the following three practical and compact experimental systems: (1) a continuous flow system containing sediment core samples; (2) a rectangular flume system; and (3) a system combining the first two. Experimental results demonstrated that SOD showed a monotonically increasing tendency as the flow velocity increased with reduction of the thickness of the diffusive boundary layer. The experimental results were compared with numerical models theoretically relating SOD and flow velocity under smooth surface conditions. The features of each experimental system are discussed. The continuous flow system is advantageous because it simultaneously produces a steady state and different dissolved oxygen (DO) conditions. The rectangular flume system is suitable for fundamental studies of hydrodynamic effects on SOD because it makes controlling hydrodynamic conditions easy, while the combined system is appropriate for studying the effect of microscopic phenomena on exchange rates, as it can reproduce natural microscopic physicochemical processes.

Introduction

When considering the material balance in a body of water (or a water column), it is important to quantify inputs and outputs, as well as internal material cycling. Material exchanges are caused by upstream inflow, downstream outflow, advection accompanying a tide, turbulent diffusion, exchange at the air-water interface, and exchange at the sediment-water interface. The relative importance of each item depends on the characteristics of the body of water. In an enclosed and eutrophic area, mass transfer at the sediment-water interface plays an important role in the material balance (Sayama 1996; Slomp et al. 1998; Utsumi et al. 1998). Dissolved oxygen (DO), the strongest biologically useful oxidant (Glud et al. 1994; Bakker and Helder 1993; Steinberger and Hondzo 1999), is especially important because of its major role in photosynthesis, respiration, and metal oxidation. Sediment oxygen demand (SOD), for example, is a dominant sink term in the DO balance (Nakamura and Stefan 1994; Steinberger and Hondzo 1999), and has a great influence on material balance and benthic organisms.
There have been many studies on SOD, and some researchers have succeeded in making quantitative estimates of in situ DO variations (e.g., Kemp and Boynton 1980; Bakker and Helder 1993). However, there are still some difficulties in assessing the SOD value, because it is strongly influenced not only by biochemical processes in the sediment, but also by hydrodynamic conditions. In studies using a benthic chamber, the overlying water is usually stirred to homogenize the overlying water and to simulate hydrodynamic conditions outside the chamber (Glud et al. 1995).
Since the 1980s, the flow velocity over the sediment has been shown to affect mass transfer rates at the sediment-water interface (Belanger 1981; Boynton et al. 1981; Jørgensen and Des Marais 1990; Nakamura and Stefan 1994; Mackenthum and Stefan 1998; Inoue et al. 2000; Arega and Lee 2005) by the formation of a diffusive boundary layer. This layer is typically less than 1 mm, and the concentrations of solutes abruptly change within it (Santschi et al. 1983; Revsbech and Jørgensen 1986; Sayama 1990; Glud et al. 1995). Such a boundary layer can act as a region for controlling scalar transport (Steinberger and Hondzo 1999) and can influence some biochemical reactions occurring just below the sediment-water interface. Some previous studies have theoretically analyzed this effect of velocity under smooth surface conditions by applying turbulent and/or laminar boundary layer theory (e.g., Nakamura and Stefan 1994; Higashino and Stefan 2005a). Higashino and Kanda (1998) studied the effect of flow velocity and sediment properties on material fluxes experimentally by using artificial sediment. A method was proposed by Berg et al. (2003) and Kuwae et al. (2006) for measuring SOD in situ without disturbing hydrodynamic conditions using an acoustic Doppler velocimeter and DO microelectrode. At present, their method is the most advanced. However, the required devices are expensive and require skilled operation. Moreover, experimental SOD estimates intending to clarify an individual process cannot be made by this method. Therefore, laboratory experiments using real sediment samples collected in the field are still important and need to be improved.
Although the effects of the hydraulic condition on SOD have been recognized and discussed, few experiments using real sediment samples collected in the field have been performed to evaluate the effect of hydrodynamic conditions on the mass transfer rate at the sediment-water interface (Brink et al. 1989; Mackenthum and Stefan 1998). One of the main reasons for this is that there is no standard method for evaluating the mass flux (Arega and Lee 2005), though many in situ and laboratory incubation methods have been proposed (e.g., Miller-Way and Twilley 1996). A sufficiently long flume is suitable for providing a fundamental understanding of hydraulic impacts on SOD but would neither be practical nor suitable for routine measurement. The estimation of in situ SOD is still important, but many experimental methods aimed at practical use do not sufficiently take hydraulic conditions into account. The purpose of this study was to examine the effect of flow velocity on the diffusive transfer rate of DO at the sediment-water interface by using three practical and compact experimental systems that could examine the dependence of SOD on flow velocity. All experiments employed real sediment samples collected in the field and were performed under smooth surface conditions. Moreover, we compared the experimental and theoretical results as suggested by Nakamura and Stefan (1994), and we also compared the experimental results among three experimental methods using the mass transfer coefficient. These discussions enable us to clarify the advantages and disadvantages of each method, and they enable us to design a measurement method for estimating SOD in the field.

Experimental Methods

Continuous Flow System with Intact Sediment Cores

The first method conducts experiments in a continuous flow system with intact sediment core samples (Fig. 1). As a stirring device was used to agitate the overlying water, the flow velocity could be controlled by changing the rotation speed of the stirring device, but the flow pattern was artificial and dissimilar to that in the field. The flow velocity distribution in the core was previously measured in detail using a laser Doppler velocimeter to obtain the relationship between the rotation speed of the stirring device and the shear velocity, calculated by fitting measured velocity profiles to a logarithmic profile.
Fig. 1. Continuous flow system that uses intact sediment cores
In this system, a downstream peristaltic pump supplies water from a feeding tank to the cores and removes it to sampling bottles at a constant discharge rate. The feeding tank, cores, peristaltic pump, DO meter (TOA, DO-25A), and sampling bottles are connected by Tygon (Courbevoie, France) tubing and polypropylene connectors. Input and output DO and nutrient concentrations are measured to evaluate the mass flux at the sediment-water interface. Most of the previous experiments with core samples have used a batch system, in which environmental conditions in the overlying water may change over time, providing a proportionally negative effect on the accuracy of the exchange rate (Miller-Way and Twilley 1996). To avoid this drawback, a continuous flow system was used to produce a steady state.
Because discharge rates are determined by the diameter of the pump tube in the continuous flow system, the hydraulic residence time can be easily controlled. As a result, DO concentrations in the cores can be regulated by changing the discharge rates. SOD is given by the following equation, assuming that the DO concentration is equal to that at the outlet under well-mixed conditions:
V(dCoutdt)=CinQCoutQSODArwV
(1)
where V=volume of overlying water in the cores; Q=discharge rate; Cin and Cout=DO concentrations at the inlet and outlet, respectively; A=surface area of the sediment; and rw is the DO consumption rate per unit volume of the overlying water. A preliminary experiment using dyes confirmed that the overlying water in cores was sufficiently homogenized within a few minutes even under the lowest rotation speed; therefore, the assumption of a well-mixed environment was considered to be correct.
Integration of Eq. (1) shows that about 95% of the volume of water in a core exchanges in 3 times the hydraulic residence time. According to Miller-Way and Twilley (1996), the time until steady state is typically less than 2 times the residence time. In a steady state, SOD can be calculated using the following equation:
SOD=QA(CinCout)rwVA
(2)
Intact sediment cores were collected at a central point of Lake Shinji, a shallow eutrophic lagoon in Japan, using Plexiglas cores (85 mm diameter, 50 cm height). The cores were immediately transported to the laboratory and placed in a water bath to maintain the water-temperature at 15°C , the temperature at the sampling site. Overlying water in the sample core was agitated using a plastic propeller positioned about 11 cm above the sediment-water interface, whose rotation speed was controlled by a speed control motor. The flow velocity in the core was controlled by changing the propeller’s rotation speed. In this case, rotation speeds were 64, 93, and 135 rpm. They were low enough to avoid resuspension of the sediment particles and resulted in mean flow velocities of 1.47, 1.82, and 3.22cms1 , respectively. In this system, “mean velocity” is the spatially averaged tangential velocity in the bulk region. The overlying water, collected near the sediment core collection site, was filtered with a glass fiber filter (Whatman, GF/C) and used as the water supply, and its DO concentration was controlled by N2 and O2 bubbling. For each experiment, two supply rates and three stirring conditions were chosen, which enabled us to perform six simultaneous experiments. Duplicate cores were prepared for each condition. One reference core, which was half filled with glass beads, was also used. After a 1-day-incubation, Experiment 1 was commenced. The DO concentration in the water supply was measured at 30 min intervals, and the concentration in each exiting water stream was measured at about 6 h intervals. After Experiment 1 was completed, the supply rates to cores 1–6 were changed, and Experiment 2 was commenced after one day had elapsed. The experimental conditions are listed in Table 1.
Table 1. Experimental Conditions in the Continuous Flow System: (a) Experiment #1; (b) Experiment #2
 Cores 1,2Cores 3,4Cores 5,6Cores 7,8Cores 9,10Cores 11,12
(a) Experiment 1
Rotation speed (rpm)64931356493135
Mean flow velocity (cms1) 1.471.823.221.471.823.22
Flow rate (mLmin1) 4.874.874.873.663.663.66
Water volume (mL)990965980940960990
Hydraulic residence time (h)3.393.303.354.284.374.51
DO concentration in bulk region (mgL1) 2.041.841.711.641.161.09
 
(b) Experiment 2
Rotation speed (rpm)64931356493135
Mean flow velocity (cms1) 1.471.823.221.471.823.22
Flow rate (mLmin1) 2.212.212.213.663.663.66
Water volume (mL)990965980940960990
Hydraulic residence time (h)7.477.287.394.284.374.51
DO concentration in bulk region (mgL1) 1.540.940.811.641.161.09

Rectangular Flume System with Recirculation

The second experimental method was a batch system that involved a gastight rectangular flume (250 cm length, 12.5 cm width, 15 cm height) in which the flow velocity could be controlled (Fig. 2). Although sediment structures are more or less disturbed, fluid motion in the field (a uniform flow, not a rotary motion) can be simulated. The middle section of the flume bed has a cavity (100 cm length, 12.5 cm width, 10 cm depth) in which the sediments taken from a eutrophic river and estuary beds are placed. In this type of flume, the sediment samples will be more or less disturbed.
Fig. 2. Rectangular flume system and its recirculation setup
Both ends of the flume are connected by pipes made of vinyl chloride (100 mm diameter), and a pump, valve, flow meter, and a water-temperature controller (ORION, LPA3) are installed between the pipes. Water in the flume is circulated through the pipes in one direction using the pump. The rotation speed of the pump and the flow velocity in the flume are regulated by a controller. In our experiment, water-temperature was also controlled ( ±0.1°C accuracy). Water in the flume was sampled from the sampling stopcock without being exposed to the air. The flow velocity profiles had been previously measured using a laser Doppler velocimeter (TSI, Model 9710), and the relationship between the cross-sectional mean flow velocity and the shear velocity was obtained based on a logarithmic profile of velocity. The DO concentration is monitored using a DO meter (TOA, DO-25A), whose probe was installed at the downstream end of the flume.
SOD is obtained by considering the rate of DO concentration change in the overlying water using the following equation:
V(dCdt)=SODArwV
(3)
where V=entire volume of recirculating water in the system; C=DO concentration in the overlying water; A=surface area of the sediment; and rw is the consumption rate in the overlying water.
The sediment used in this experiment was collected from a river bed using a shovel and was immediately transported to the laboratory. The sediment contained benthic organisms such as river crabs and nereimorpha. The sediment was placed in the cavity of the flume bed and fresh water was carefully poured into the flume so as not to disturb the sediment surface. Next, the pump was started to produce a constant flow in the flume, and the water-temperature controller was set to 30°C . The experiment was started after as many of the benthic organisms had been removed as possible using tweezers. The DO concentration in the overlying water was monitored every hour during the experiment using a DO meter. Water sampling was carried out at the beginning and end of each experiment. The experiment was performed in the dark with a thick black curtain covering the entire flume to suppress photosynthesis generating DO. Three flow velocities, 1.47, 3.93, and 6.09cms1 , were used. After each experiment, air bubbles were passed through the overlying water in the flume to increase the DO concentration for the next experiment. The DO meter was calibrated with a 2-point-method that used the DO concentration measured by the Winkler method at the starting and finishing time of each experiment.

Combined System of a Rectangular Flume and Undisturbed Sediment Core

The third experimental method was a combination of the first two methods (Fig. 3). The method was originally designed by Sayama (1990). A cylindrical core with an inner diameter of 8 cm is screwed beneath the floor of the rectangular main flume (19 cm length, 12 cm width, 5 cm height), and the sediment surface is adjusted to the level of the flume surface. Water in the main flume is recirculated, and its flow velocity can be controlled by a variable pump. This system is advantageous because it can simulate natural flow conditions and use intact sediment cores. Before the experiment, the relationship between the vertical velocity profile and the rotation speed of the variable pump was obtained using a laser Doppler velocimeter (TSI, Model 9710). In addition, the relationship between the rotation speed and shear velocity was also obtained by fitting measured velocity profiles to a logarithmic profile. The feeding water inlet and the overlying water outlet are mounted on the main flume, making continuous flow experiments possible.
Fig. 3. Combined system with a rectangular flume plus an undisturbed sediment core
In this experiment, precise distributions of the DO concentration were measured in the vicinity of the sediment surface using a DO microelectrode. In the 1980s, a DO microelectrode with a high spatial resolution was developed, enabling measurement of precise DO profiles near the sediment-water interface and SOD estimation from concentration gradients (e.g., Jørgensen and Revsbech 1985). The DO microelectrode makes it possible to observe directly the effects of flow velocity on the DO profile and SOD. The glass-coated electrode used in the measurement has a platinum tip with a 5μm diameter and a very high spatial resolution. In this microsensor, the cathode is situated behind an electrically insulating membrane of silicone rubber, which is extremely permeable to oxygen (Jørgensen and Revsbech 1985). The platinum wire is covered with an inner glass pipe excluding the tip and is bathed in an electrolyte solution of KCl into which a Ag/AgCl reference electrode is immersed. The electrode is attached to a micromanipulator whose precision is 1μm , so that measurement of detailed DO profiles near the sediment-water interface is possible. A control device that automatically operates the micromanipulator enables movement of the microelectrode at different time and vertical space intervals. The micromanipulator is also attached to another traverse device that can move the microelectrode horizontally with a precision of 100μm . These two traverse devices make it possible to precisely reach any measurement point. The output from the microelectrode signals is amplified by a picoammeter and recorded.
Intact sediment cores were collected with Plexiglas cores (8 cm diameter, 30 cm height) from Lake Shinji. Water-temperature and DO concentration at the sampling location were measured. An oxic brown layer was observed at the top 2 or 3 mm of the sediment, and an anoxic layer was observed underneath. Sediment cores were immediately transported to the laboratory and placed in the experimental device as described above. The overlying water collected at the sediment core sampling location was filtered using a glass fiber filter (Whatman, GF/C), and carefully poured into the main flume, so as not to disturb the sediment surface. The silty sediment collected from Lake Shinji had a distinct sediment-water interface, allowing easy identification of the position of the sensor tip relative to the interface using a stereoscopic microscope. Since the DO concentration is directly proportional to the electrode’s output signal, the DO electrode calibration was done in the bulk region of the overlying water and the anoxic part of the sediment (Bakker and Helder 1993; Glud et al. 1994).
Vertical DO concentration profiles near the sediment-water interface were measured for various flow velocities and locations. Before making the main measurements, we checked the horizontal DO homogeneity to find out a representative measurement point. Given our premeasurement results, we used profiles measured on the downstream side of the core, as discussed below. Moreover, since Inoue et al. (2000) predicted that SOD values would reach almost steady values within 20 min, DO profiles were measured more than 30 min after velocity changes.

Common Measurement Items

The DO consumption rate per unit volume of the overlying water, rw , was measured by a quasi-BOD method at the experimental temperature. The porosity and sediment oxygen consumption rate per unit volume were also measured immediately after each experiment. Known volumes of particular layers of sediments were removed in narrow sediment cores using end-cut syringes from the sediment employed in each experiment. Some portions of the sediment were dried at 60°C for 2 days, and their porosities were estimated from the weight difference before and after the drying process. Other portions were used to measure the volumetric oxygen consumption rate of the sediment, rs , which was determined by the method of Hosoi et al. (1992); the DO concentration of water clouded with sediment was measured. In the following analysis, rs values obtained during the first half day were employed to avoid negative effects due to a decrease in the sediment organic-matter content after a half day. This parameter will be discussed below.

Results and Discussion

Experimental Results

From the results of the continuous flow system, it is evident that the DO concentration decreases when the flow velocity increases (data not shown). This is probably caused by a decrease in the thickness of the diffusive boundary layer and an increase in the DO concentration gradient at the sediment-water interface as the flow velocity increases (e.g., Boudreau 2001). The relationship between SOD calculated from Eq. (2) and the rotation speed is given in Fig. 4. In Fig. 4, sediment cores are divided into two groups, namely Core 1 to Core 6 and Core 7 to Core 12, because cores in the same group experienced almost the same experimental conditions except for rotation speed. Each experiment continued for about 30 h. Since the DO concentration in each water outlet stream was measured at about 6 h intervals, we calculated SOD for 5 times. As the experiments were performed in duplicate, standard deviations were calculated from the 10 values. It can be seen that SOD increases when the rotation speed (flow velocity) increases. In addition, SOD increases with an increasing supply rate because the DO concentration in the overlying water strongly depends on the supply rate (i.e., turnover time). Therefore, the DO concentration gradient at the sediment-water interface might be less under low supply rate conditions.
Fig. 4. SOD versus rotation speed for the continuous flow system with intact sediment cores. Error bars show standard deviations.
In the rectangular flume system, the DO concentration decreases with time because of DO consumption in the sediment. However, the decreasing rates of the DO concentration varied depending on the flow velocities. Fig. 5 shows that high flow velocity conditions corresponded to high SOD, and that SOD showed a monotonically increasing tendency as the DO concentration increased, except at 1.47cms1 . These results qualitatively reveal the same tendency as results obtained from the continuous flow system and indicate that the limiting process for SOD in these experiments could be considered the diffusive transport of DO in the boundary layer around the sediment-water interface (e.g., Nakamura and Stefan 1994).
Fig. 5. Relationship between DO concentration and SOD in the rectangular flume system
Fig. 6 shows vertical DO microprofiles measured in the combined system around the diffusive boundary layer with different flow velocities. Theoretically calculated DO profiles discussed in the following analysis are also presented. Flow velocities changed from 0.24 up to 10.1cms1 , where no visible sediment resuspension was observed. When the flow velocity was 10.1cms1 , however, it was visibly confirmed that some sediment particles were rolling on the sediment surface as a kind of bed load. The figure clearly shows that as flow velocity increases, the thickness of the diffusive boundary layer, δ , decreases (0.62 mm for 0.24cms1 and 0.28 mm for 10.1cms1 ), the DO concentration at the sediment-water interface increases ( 4.1mgL1 for 0.24cms1 and 6.0mgL1 for 10.1cms1 ), and the oxygen penetration depth increases (2.0 mm for 0.24cms1 and 2.9 mm for 10.1cms1 ). Reduction of the diffusive boundary layer thickness leads to an increased concentration gradient in the layer, which indicates an increase in SOD.
Fig. 6. Velocity dependence of the DO profile around the diffusive boundary layer

Effects of Flow Velocity on SOD

Nakamura and Stefan (1994) expressed the theoretical relationship between SOD and flow velocity under smooth surface conditions when the volumetric rate of oxygen consumption is constant as follows:
SOD=1+(1+U2)1/2U
(4)
where SOD and U=nondimensional SOD and flow velocity, respectively, as defined by
SOD=SOD2DsrsCO(z=)
(5)
U=2nπCfSc3/4U2CO(z=)Dsrs
(6)
where Ds=apparent diffusion coefficient of DO in the sediment; rs=DO consumption rate per unit volume of the sediment; CO(z=)=DO concentration in the bulk region in the overlying water; n=numerical constant ( =0.124 , Deissler 1955); Cf=friction coefficient; Sc=Schmidt number; and U=mean flow velocity.
Fig. 7 shows the relationship between nondimensional SOD and velocity as defined by Eq. (4). The experimental data obtained using the continuous flow system were scattered. Although the theoretical curve shows an increasing tendency of SOD as U increases, theoretical SOD values were considerably lower than the experimental values. As a large number of spiomorpha were observed in the sediment cores, the effect of their respiration and bioturbation on SOD may have been substantial (Andersen and Helder 1987; Bakker and Helder 1993). However, there was no clear surface area increase due to the existence of benthic organisms. In addition, the estimation of U might be open to further discussion, because estimating U from the measured data of flow velocity distribution in the core was difficult. Fig. 8 compares theoretical values with experimental values obtained from the rectangular flume system. The theoretical values qualitatively reproduced the experimental values. Fig. 8 shows a better estimation than Fig. 7. The theoretical predictions show that hydrodynamic control is significant when nondimensional velocity is less than 5. In this region, diffusive transfer in the boundary layer controls the total flux. The closer agreement between theoretical and experimental values in Fig. 8 shows that the diffusive transport model is reliable for studying the boundary layer. However, the scatter in the experimental results in Fig. 8 shows the need for further refining the oxygen kinetics model in the sediment. As shear stress in the developing boundary layer is expected to be larger than that in the fully developed boundary layer, experimental SOD values should be larger than the theoretical values. This might cause the discrepancy shown in Fig. 8.
Fig. 7. Comparison of theoretical values with experimental values obtained from the continuous flow system
Fig. 8. Comparison of theoretical values with experimental values obtained from the rectangular flume system
Figs. 7 and 8 show that all SOD values calculated from the experimental results exceed the theoretical estimates, by factors of 2–3 in Fig. 7 and 1–1.5 in Fig. 8. This finding agrees with the report of Arega and Lee (2005) that SOD calculated from the DO profile by Fick’s law of diffusion was smaller than the value obtained from continuous flow and batch operations. In the continuous flow experiment, the discrepancy shown in Fig. 7 was considered to be due to the respiration of a large number of spiomorpha observed on the sediment surface (Rasmussen and Jørgensen 1992). Calculating SOD only from the concentration gradient at the sediment-water interface might be inappropriate when the macrofaunal effect is high. Another possible cause of the discrepancy was the higher effective surface area, as pointed out by Gundersen and Jørgensen (1990). Such discrepancies have been also reported for metabolism rates of coral reef algal turfs (Carpenter and Williams 2007). Precise measurements are necessary for further refinement of these results, using, for example, oxygen microelectrodes that have a high resolution in space (e.g., Jørgensen and Des Marais 1990).

Microprofile of DO Concentration

DO concentration profiles are calculated using a diffusion equation. Assuming that: (1) there is a steady state; (2) the sediment surface is a fixed, smooth, and flat bed; (3) velocity and DO concentration are homogeneous except in the vertical direction; (4) the boundary layer is thin enough for its DO consumption to be negligible and shear stress in it to be approximated as a constant, according to Fick’s second law of diffusion, the diffusion equation of DO can be expressed as follows:
dCO(z)dt=Dzd2CO(z)dz2(inthewater)
(7)
dCO(z)dt=εDsd2CO(z)dz2rs(inthesediment)
(8)
where Dz=diffusion coefficient of DO in the overlying water and ε=porosity . Dz is expressed as follows (Deissler 1955), assuming that the turbulent Schmidt number is unity:
Dz(z+)ν=n2u+z+{1exp(n2u+z+)}
(9)
where ν=kinematic viscosity u+ and z+=nondimensional velocity and vertical coordinate, defined by
u+=uu
(10)
z+=zuν
(11)
where u=flow velocity and u=shear velocity. On the other hand, Ds is expressed as follows (Berner 1980):
Ds=Dmθ2=DmεF=εm1Dm
(12)
where Dm=molecular diffusion coefficient of DO; θ=tortuosity ; and m=numerical constant ( =1.8 , Berner 1980). Eq. (7) represents the DO flux across the diffusive boundary layer, whereas Eq. (8) represents the DO flux at the sediment-water interface. Assuming that the diffusive boundary layer is thin enough for its DO consumption to be negligible, the two fluxes must coincide. Therefore, Eqs. (7) - (8) are combined by eliminating the concentration at the interface.
A comparison of the theoretical DO concentration profile with the measured profile of the present study is shown in Fig. 6. The theoretical calculations can qualitatively reproduce the decrease in diffusive boundary layer thickness and the increase of oxygen penetration depth with the increase of flow velocity. However, the estimate under the bed load condition was unsatisfactory. Some phenomenon such as pore water advection may have occurred near the sediment-water interface to cause this result.
In order to estimate SOD the concentration gradient in the diffusive boundary layer must be precisely known. A method of estimating SOD by multiplying the DO gradient and molecular diffusivity has been considered to be reliable as long as the diffusive boundary layer thickness is larger than the depth resolution. For a Schmidt number, Sc, of 500, typical for DO, the diffusive boundary layer thickness would be 2n/u . Therefore, the DO concentration should be measured at intervals of much less than 2n/u to obtain an accurate estimate of SOD. In that case, the DO concentration gradient could be reproduced accurately. However, quantitative estimation of total SOD is still difficult even without benthic fauna. One possible explanation of this is that the surface area was underestimated because the microscopic topography at the sediment surface was neglected.

Mass Transfer Coefficient

As nondimensional parameters such as U and SOD are not commonly used, comparing these values with results of other papers is difficult. Therefore, we introduce and discuss the mass transfer coefficient, Kc . Kc is basically defined by Eq. (13)
Kc=Dmδ
(13)
where δ=effective diffusive boundary layer thickness, obtained from extrapolating the linear gradient at the sediment-water interface to the bulk DO concentration (Jørgensen and Revsbech 1985). Kc was calculated using Eq. (13) and the experimental results for the combined system, because δ could be directly measured in this system. However, δ could not be directly measured in the other two systems, so Kc was calculated using Eq. (14), which was mathematically derived from Eq. (13) for the continuous flow system, the rectangular flume system, and continuous flow measurements in the combination system
Kc=SODCO(z=)CO(z=0)
(14)
where CO(z=0)=DO concentration at the sediment-water interface. In the process of the model formulation in Nakamura and Stefan (1992), CO(z=0) was expressed using Eq. (15)
CO(z=0)=1+U2+11+U2+1CO(z=)
(15)
Arega and Lee (2005) showed that Kc is related to the shear velocity, u , and the Schmidt number, SC, as,
Kc=cuSc2/3
(16)
where c=numerical coefficient.
Fig. 9 shows the relationship between Kc and uSc2/3 . These Kc values are of the same order as the 0.0139mms1 of Steinberger and Hondzo (1999) and the 0.0020.013mms1 reported by Arega and Lee (2005). Fig. 9 allows us to calculate the coefficient for each system, giving cCFS=0.179 for the continuous flow system, cRFS=0.090 for the rectangular flume system, cCSE=0.089 for DO microelectrode measurements in the combination system, and cCSC=0.326 for continuous flow measurements in the combined system. These c values are larger than the value of 0.0645 obtained from empirical coefficients by Nakamura and Mikogami (1994). It is worth noting that cCSC is much larger than cCSE by a factor of 3.5. Arega and Lee (2005) carried out an experiment using a cylindrical SOD chamber, and they also obtained different coefficients of c=0.199 for the total SOD and c=0.106 for the diffusive SOD. According to cRFS and cCSE in this paper and c for the diffusive SOD in Arega and Lee (2005), c values under a diffusion-oriented condition seem to be nearly 0.1. However, other experiments show larger values, and the differences may be caused by processes other than diffusion, such as bioturbation and/or bioirrigation. Rasheed et al. (2006) also pointed out the same tendency of release rates for inorganic nitrogen, phosphate, and silicate from the sediment. Further studies are required to clarify the reason for such a discrepancy.
Fig. 9. Relationship between Kc and uSc2/3 , where CFS, RFS, CSE, and CSC represent the continuous flow system, the rectangular flume system, the DO microelectrode measurements in the combined system, and the continuous flow measurements in the combination system, respectively

Features of Each Method

The first method was used for experiments in a continuous flow system with sediment core samples. A stirring device was employed to agitate the overlying water. The advantage of this system is that it can simultaneously produce a steady state and multiple environmental conditions. As controlling water-quality conditions for each core is simple with this system, it is suitable for reproducing field experiments with multiple samples. As can be seen clearly from Eq. (1), DO concentrations in the overlying water can be controlled by changing the discharge rate. This is advantageous for studying the anaerobic release of chemicals by controlling DO and nitrate concentrations, which are responsible for the redox potential. In order to evaluate the actual mass flux in the environment, intact sediment cores should be used, but undisturbed sediment is not always necessary for fundamental studies on the effects of hydrodynamics and water-quality on fluxes at the sediment-water interface. However, the hydraulic condition in the cores is artificial.
The second method was performed with a rectangular, long flume in which the flow velocity was controlled. It is suitable for fundamental study of processes, such as hydrodynamic control of mass transfer, rather than assessment of the actual environmental flux. For instance, rectangular roughness elements can be placed on the flume floor to investigate the effects of surface roughness in a long flume in which the turbulent boundary layer is fully developed downstream. In practice, however, shorter flumes are usually used to reduce cost. Experimental results using them could be influenced by nonuniform shear stresses and shear velocity in the downstream direction. Higashino and Stefan (2005b) presented a model of SOD for a sediment bed of finite length. According to their model predictions, a sediment bed of 10,000ν/u length in the downstream direction is needed to make the effect of diffusive entry length on the mean SOD negligible. u is the shear velocity, calculated by Eq. (17)
u=Ug1/2lR1/6
(17)
where g=acceleration gravity; l=Manning ’s roughness coefficient ( =0.024 for fine gravel, Arcement and Schneider 1989); and R=hydraulic radius. From this point of view, the sediment bed lengths required were about 436 cm for U=1.47cms1 , about 163 cm for U=3.93cms1 , and about 105 cm for U=6.09cms1 . Since the sediment bed length was 100 cm in this study, experiments under high flow velocity conditions were considered to be acceptable. However, under low flow velocity conditions, the experimental results should be treated carefully and could be improved.
The third method consisted of a combined system of the aforementioned two methods: a cylindrical, intact core sample was attached beneath the main rectangular flume. This system is advantageous because it can simulate natural flow conditions and use intact sediment cores. It also is more appropriate for studying actual exchange rates, as it can reproduce natural physicochemical processes not only in the sediment but also in the overlying water.
Consequently, the rectangular flume system and microprobe measurements in the combined system are suitable for fundamental studies of the diffusion process around the sediment-water interface, especially with respect to hydrodynamic control of diffusive transfer rate. On the other hand, the continuous flow system and the continuous flow measurement in the combined system are advantageous for measuring actual exchange rates including other processes. In practice, a system should be appropriately chosen by considering the condition of collected sediment, the number of samples, the aim of the work, and so on. For example, if one intends to measure microprofiles around the sediment-water interface, one should use a system with a microelectrode. If hydrodynamic control is the main object, we recommend using a rectangular flume. On the contrary, intact sediment cores should be prepared for biochemical assessment of mass fluxes in the field, because biochemical conditions in the surface layer of the sediment are easily disturbed, and it is quite important to preserve vertical distribution of particles and solutes. In addition, for multiple measurements, a system such as the continuous flow system introduced this paper is convenient.

Summary and Conclusions

The effect of flow velocity on the diffusive transfer rate of DO at the sediment-water interface was quantitatively investigated with three experimental methods and real sediment samples collected in the field. The first experimental method consisted of a continuous flow system with intact core samples of the sediment. In this experiment, the flow velocity in the overlying water was controlled by changing the rotation speed of the stirring device, but the flow pattern was artificial and dissimilar to that in the field. Experimental results from the continuous flow system showed that SOD increases when the flow velocity increases and also when the DO concentration in the overlying water increases. The second method involved a rectangular flume in which flow velocity could be controlled. In this experiment, sediment structures were more or less disturbed, but fluid motion in the field could be simulated. This experiment showed that the DO concentration in the overlying water decreases with time, and decreasing rates of DO concentrations show a monotonically increasing tendency as flow velocities increase. The third experimental method combined the two previous methods. In this experiment, the precise distributions of DO were measured in the vicinity of the sediment surface by using a DO microelectrode, and a continuous flow experiment was conducted simultaneously. The experimental results showed that as flow velocity increases, the thickness of the diffusive boundary layer decreases, and the oxygen penetration depth increases. Furthermore, SOD values calculated by continuous flow measurements in the combined system are much larger than those calculated from the DO profile by a factor of 3.5. The numerical prediction of Nakamura and Stefan (1994) qualitatively reproduced experimental values, but predicted values were smaller than observed. Calculating SOD only from the concentration gradient at the sediment-water interface might be difficult. Further studies are required to elucidate such discrepancies and determine the best method for estimating the “actual” SOD.

Notation

The following symbols are used in this paper:
A
=
surface area of the sediment (L2) ;
C
=
material concentration in the overlying water (ML3) ;
Cf
=
friction coefficient (−);
Cint
=
concentration at the inlet (ML3) ;
CO(z=0)
=
DO concentration at the sediment-water interface (ML3) ;
CO(z=)
=
DO concentration in the bulk region in the overlying water (ML3) ;
Cout
=
concentration at the outlet (ML3) ;
c
=
numerical coefficient (−);
Dm
=
molecular diffusion coefficient of DO (L2T1) ;
Ds
=
apparent diffusion coefficient of DO in the sediment (L2T1) ;
Dz
=
diffusion coefficient of DO in the overlying water (L2T1) ;
F
=
mass flux at the sediment-water interface (ML2T1) ;
g
=
acceleration gravity (LT2) ;
Kc
=
mass transfer coefficient (−);
l
=
Manning’s roughness coefficient (−);
m
=
numerical constant (−);
n
=
numerical constant (−);
Q
=
discharge rate (L3T1) ;
R
=
hydraulic radius (L);
rs
=
DO consumption rate per unit volume of the sediment (ML3T1) ;
rw
=
DO consumption rate per unit volume of the overlying water (ML3T1) ;
Sc
=
Schmidt number;
U
=
mean flow velocity (LT1) ;
U
=
nondimensional mean flow velocity (−);
u
=
flow velocity (LT1) ;
u+
=
nondimensional velocity (−);
u
=
shear velocity (LT1) ;
V
=
volume of overlying water (L3) ;
z+
=
nondimensional velocity (−);
δ
=
diffusive boundary layer thickness (L);
ε
=
porosity;
θ
=
tortuosity; and
ν
=
kinematic viscosity (L2T1) .

Acknowledgments

We are grateful to Dr. M. Yamamuro, Dr. Y. Ishitobi, Dr. M. Sayama and Dr. H. Kamiya for many valuable discussions. We also thank the students of the Dept. of Maritime Systems Engineering, Faculty of Engineering, Kyushu Univ. for their technical help. The manuscript was greatly improved by valuable comments from anonymous reviewers.

References

Andersen, F. O., and Helder, W. (1987). “Comparison of oxygen microgradients, oxygen flux rates and electron transport system activity in coastal marine sediments.” Mar. Ecol.: Prog. Ser., 37, 259–264.
Arcement, G. J., Jr., and Schneider, V. R. (1989). “Guide for selecting Manning's roughness coefficients for natural channels and flood plains.” United States Geological Survey Water-supply Rep. No. 2339, Metric Version.
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.
Belanger, B. T. (1981). “Benthic oxygen demand in Lake Apopka, Florida.” Water Res., 15(2), 267–274.
Berg, P., Roy, H., Janssen, F., Meyer, V., Jørgensen, B. B., Huettel, M., and de Beer, D. (2003). “Oxygen uptake by aquatic sediments measured with a novel non-invasive eddy-correlation technique.” Mar. Ecol.: Prog. Ser., 261, 75–83.
Berner, R. A. (1980). “Diagenetic physical and biological processes.” Early diagenesis, Princeton University Press, Princeton, N.J., 15–56.
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.
Boynton, W. R., Kemp, W. M., Osborne, C. G., Kaumeyer, K. R., and Jenkins, M. C. (1981). “Influence of water circulation rate on in situ measurements of benthic community respiration.” Mar. Biol. (Berlin), 65(2), 185–190.
Brink, M. R. B., Gust, G., and Chavis, D. (1989). “Calibration and performance of a stirred benthic chamber.” Deep-Sea Res., 36(7), 1083–1101.
Carpenter, R. C., and Williams, S. L. (2007). “Mass transfer limitation of photosynthesis of coral reef algal turfs.” Mar. Biol. (Berlin), 151(2), 435–450.
Deissler, R. G. (1955). “Analysis of turbulent heat transfer, mass transfer, and friction in smooth tubes at high Prandtl and Schmidt numbers.” National Advisory Committee for Aeronautics Rep. No. 1210, Lewis Flight Propulsion Lab, Cleveland, 69–82.
Glud, R. N., Gundersen, J. K., Barker, B., 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.
Glud, R. N., Gundersen, J. K., Revsbech, N. P., Jørgensen, B. B., and Hüttel, M. (1995). “Calibration and performance of the stirred flux chamber from the benthic Lander Elinor.” Deep-Sea Res., 42(6), 1029–1042.
Gundersen, J. K., and Jørgensen, B. B. (1990). “Microstructure of diffusive boundary layers and the oxygen uptake of the sea floor.” Nature, 345(6276), 604–607.
Higashino, M., and Kanda, T. (1998). “Diffusional mass transfer from bottom sediment to flowing water.” Annu. J. Hydrau. Eng. JSCE, 42, 745–750.
Higashino, M., and Stefan, H. G. (2005a). “Sedimentary microbial oxygen demand for laminar flow over a sediment bed of finite length.” Water Res., 39(14), 3153–3166.
Higashino, M., and Stefan, H. G. (2005b). “Oxygen demand by a sediment bed of finite length.” J. Environ. Eng., 131(3), 350–358.
Hosoi, Y., Murakami, H., and Kozuki, Y. (1992). “Oxygen consumption by bottom sediment.” Proc. JSCE, Hydrau. Sanitary Eng., 456 (in Japanese with English abstract).
Inoue, T., Nakamura, Y., and Adachi, Y. (2000). “Non-steady variations of SOD and phosphate release rate due to changes in water quality of the overlying water.” Water Sci. Technol., 42(3–4), 265–272.
Jørgensen, B. B., and Des Marais, D. J. (1990). “The diffusive boundary layer of sediments: Oxygen microgradients over a microbial mat.” Limnol. Oceanogr., 35(6), 1343–1355.
Jørgensen, B. B., and Revsbech, N. P. (1985). “Diffusive boundary layers and the oxygen uptake of sediments and detritus.” Limnol. Oceanogr., 30(1), 111–121.
Kemp, W. M., and Boynton, W. R. (1980). “Influence of biological and physical processes on dissolved oxygen dynamics in an estuarine system: Implications for measurement of community metabolism.” Estuarine Coastal Shelf Sci., 11(4), 407–431.
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.
Mackenthun, A. A., and Stefan, H. G. (1998). “Effect of flow velocity on sediment oxygen demand: Experiments.” J. Environ. Eng., 124(3), 222–229.
Miller-Way, T., and Twilley, R. R. (1996). “Theory and operation of continuous flow systems for the study of benthic-pelagic coupling.” Mar. Ecol.: Prog. Ser., 140, 257–269.
Nakamura, Y., and Mikogami, M. (1994). “Near-bed turbulence and mass transfer at the sediment-water interface.” Annu. J. Hydrau. Eng. JSCE, 38, 223–228 (in Japanese with English abstract).
Nakamura, Y., and Stefan, H. G. (1992). “Sediment oxygen demand in Lakes: Dependence on near-bottom flow velocities.” Rep. No. 335, St. Anthony Falls Hydraulic Laboratory, University of Minnesota, Minn.
Nakamura, Y., and Stefan, H. G. (1994). “Effect of flow velocity on sediment oxygen demand: Theory.” J. Environ. Eng., 120(5), 996–1016.
Rasheed, M., Al-Rousan, S., Manasrah, R., and Al-Horani, F. (2006). “Nutrient fluxes from deep sediment support nutrient budget in the oligotrophic waters of the Gulf of Aqaba.” J. Oceanogr., 62(1), 83–89.
Rasmussen, H., and Jørgensen, B. B. (1992). “Microelectrode studies of seasonal oxygen uptake in coastal sediments: Role of molecular diffusion.” Mar. Ecol.: Prog. Ser., 81, 289–303.
Revsbech, N. P., and Jørgensen, B. B. (1986). “Microelectrodes: Their use in microbial ecology.” Advances in microbial ecology, Vol. 9, Plenum, New York, 293–352.
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.
Sayama, M. (1990). “A method to measure oxygen microprofile in the surface sediment.” J. Adv. Mar., 2, 63–71 (in Japanese with English abstract).
Sayama, M. (1996). “Nitrogen cycling process and its release at the sediment surface in coastal area.” J. NIRE., 5, 291–306 (in Japanese with English abstract).
Slomp, C. P., Malschaert, J. F. P., and Van Raaphorst, W. (1998). “The role of adsorption in sediment-water exchange of phosphate in North Sea continental margin sediment.” Limnol. Oceanogr., 43(5), 832–846.
Steinberger, N., and Hondzo, M. (1999). “Diffusional mass transfer at sediment-water interface.” J. Environ. Eng., 125(2), 192–200.
Utsumi, M., Nojiri, Y., Ytow, N., and Seki, H. (1998). “Dynamics of attached bacteria at the water-sediment interface in a mesotrophic swampy bog of Japan.” J. Oceanogr., 54(2), 179–184.

Information & Authors

Information

Published In

Go to Journal of Environmental Engineering
Journal of Environmental Engineering
Volume 135Issue 11November 2009
Pages: 1161 - 1170

History

Received: May 29, 2007
Accepted: Jun 18, 2009
Published online: Oct 15, 2009
Published in print: Nov 2009

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]
Executive 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