Open access
Technical Papers
Dec 15, 2017

CFD Methodology to Determine the Hydrodynamic Roughness of a Surface with Application to Viscous Oil Coatings

Publication: Journal of Hydraulic Engineering
Volume 144, Issue 2

Abstract

Water-lubricated pipe flow technology is an economic alternative for the long-distance transportation of viscous oils, such as heavy oil and bitumen. In the industrial-scale application of this technology, a thin oil film is always observed to coat the pipe wall. The natural process of wall coating during the lubrication is often referred to as wall fouling. A wall-fouling layer produces ultrahigh values of hydrodynamic roughness (1  mm), which have not been studied sufficiently to date. In this work, the hydrodynamic effects of a viscous wall-coating layer were experimentally investigated. A customized flow cell was used for the purpose. The equivalent sand grain (hydrodynamic) roughness was determined using a methodology involving computational fluid dynamics (CFD) simulations. The hydrodynamic roughness was also determined from the measured topology (physical roughness) of the surface. Additional verification of the method was obtained by applying it to analyze the hydrodynamic roughness produced by sandpapers and biofouling layers. The primary outcome of the present study is the validation and application of a CFD-based methodology to quantify the hydrodynamic roughness produced by any surface, including viscous oil coatings and biofouled surfaces. Additionally, it has been shown that the hydrodynamic roughness of a viscous oil coating, for the range of conditions tested here, is much more dependent on the coating thickness than on the Reynolds number. This has significant implications for the modeling of lubricated pipeline flows involving heavy oil and water.

Introduction

Known reserves of nonconventional oils represent one of the largest global petroleum resources (Nunez et al. 1998; Saniere et al. 2004; CAPP 2015). These reserves mostly comprise asphaltic, dense, and very viscous oils, such as bitumen and heavy oil. The production and transportation of these nonconventional oils require extraordinary techniques that are not needed for conventional petroleum deposits. Various mining and in situ production technologies are used to extract these nonconventional oils. After extraction, these viscous oils are transported from a company’s various production sites to its upgrading facilities (Joseph et al. 1999; Sanders et al. 2004). Pipelines are most commonly used for this transportation, provided the effective viscosity of the oil can be reduced through heating, dilution, partial upgrading, or emulsification (Nunez et al. 1998; Saniere et al. 2004; Jean et al. 2005). These methods require substantial capital investments and operating costs, which contribute significantly to the higher production costs of heavy oil. As a result, the oil industry is keen to use more cost-effective technology for heavy oil transportation (Nunez et al. 1998; Guevara et al. 1998; Saniere et al. 2004; Hart 2014).
Water-assisted pipeline transportation is a promising alternative technology for transporting viscous oils like heavy oil and bitumen. Here, viscous oil flows in the core and water flows through the annulus. The annular water film mostly protects the viscous oil from touching the pipe wall and thereby acts as a lubricant. The lubricating water is either injected or already present in the mixture (Arney et al. 1993; Joseph et al. 1999; McKibben et al. 2000; Sanders et al. 2004; M. McKibben, R. Gillies, S. Rushd, and S. Sanders, “A new model to predict pressure losses in the water-assisted pipeline transport of bitumen or heavy oil,” submitted, J. Petrol Explor. Prod. Technol., Springer, Berlin). This pipeline transportation technology is referred to as lubricated pipe flow (LPF). It requires much lower energy input compared with the transportation of viscous oil alone in the pipeline (Rodriguez et al. 2009; Crivelaro et al. 2009).
A concern for the application of LPF is that some oil tends to permanently adhere to the pipe wall (Saniere et al. 2004). This phenomenon is called wall fouling. The primary causes of wall fouling appear to be through direct contact of the oil core with the wall generally because of a low water cut (i.e., a thin lubricating layer), low operating velocities such that the hydrodynamic lift forces are not strong enough to keep the core away from the wall, and density differences between the core and the annular fluids (Arney et al. 1996; McKibben et al. 2000; Schaan et al. 2002; Rodriguez et al. 2009). Other contributors to wall fouling include pipe material (wettability), pipe shutdowns (i.e., no-flow conditions), and transport of oil droplets from the annulus to the wall (Arney et al. 1993; Joseph et al. 1999; Sanders et al. 2004). Even though frictional pressure loss in a fouled pipe is higher compared with that for similar transportation in an unfouled pipe, the loss is substantially lower than what would be expected for transporting only heavy oil or bitumen (Arney et al. 1996; Joseph et al. 1999; McKibben et al. 2000; Schaan et al. 2002; Sanders et al. 2004; Rodriguez et al. 2009; Crivelaro et al. 2009; M. McKibben, R. Gillies, S. Rushd, and S. Sanders, “A new model to predict pressure losses in the water-assisted pipeline transport of bitumen or heavy oil,” submitted, J. Petrol Explor. Prod. Technol., Springer, Berlin). It should be noted that wall fouling does not appear to destabilize the annular (lubricated) flow regime even though it produces higher friction losses (Joseph et al. 1999; Rodriguez et al. 2009; M. McKibben, R. Gillies, S. Rushd, and S. Sanders, “A new model to predict pressure losses in the water-assisted pipeline transport of bitumen or heavy oil,” submitted, J. Petrol Explor. Prod. Technol., Springer, Berlin).
Formation of a wall-fouling layer of oil is practically unavoidable in the industrial-scale applications of LPF technology (McKibben et al. 2000; Schaan et al. 2002; Shook et al. 2002; Rodriguez et al. 2009). Different degrees of wall fouling are experienced in various applications of LPF, making it possible to divide LPF into two major categories depending on the extent of fouling: core annular flow (CAF) and continuous water-assisted flow (CWAF). Core annular flow primarily denotes a somewhat idealized concept of LPF because it involves a core of viscous oil lubricated with a water annulus in an unfouled pipe (Arney et al. 1993). In practice, most (if not all) commercial applications of LPF can be categorized as CWAF.
Presently, a reliable model to predict pressure losses in CWAF is not available (Shook et al. 2002; Hart 2014; M. McKibben, R. Gillies, S. Rushd, and S. Sanders, “A new model to predict pressure losses in the water-assisted pipeline transport of bitumen or heavy oil,” submitted, J. Petrol Explor. Prod. Technol., Springer, Berlin). Although numerous empirical, semimechanistic, and numerical models have been proposed (e.g., Arney et al. 1993; Ho and Li 1994; Joseph et al. 1999; Rodriguez et al. 2009; Crivelaro et al. 2009), all are very limited in applicability. Some are only appropriate for CAF, and others are highly system-specific. None of the existing models explicitly addresses the effect of wall fouling on frictional pressure losses.
Although the layer is relatively thin (compared with the pipe diameter), it is textured or rippled, which can significantly increase hydrodynamic roughness (Brauer 1963; Picologlou et al. 1980; Shook et al. 2002). The rippling effect (i.e., the irregular waves that form on the viscous coating layer) is associated with interfacial instability due to viscosity stratification (Hooper and Boyd 1987; Kushnir et al. 2014). Such instability is a result of the interaction of the two fluids through the velocity and viscous stresses at the interface (Tilley et al. 1994; Govindarajan and Sahu 2014). The instability effect is discussed extensively in the literature, but it is not the focus of this study; instead, the goal here is to evaluate the effect of the rippled coating on hydrodynamic roughness.
The engineering scale for hydrodynamic roughness is the Nikuradse sand grain equivalent (Flack and Schultz 2010). This kind of equivalent roughness is used extensively for commercial metal pipes or channels. Similar roughness is also used for various atypical rough walls, such as metal walls with nonuniform roughness, minichannels, or biofilms on a solid wall (Picologlou et al. 1980; Bayazit 2010; Kandlikar et al. 2005). A biofilm is somewhat similar to a viscous oil coating in that it is conformable and can substantially increase hydrodynamic roughness (Picologlou et al. 1980; Barton et al. 2005; Schultz 2007; Lambert et al. 2009; Andrewartha 2010). Although some measurements of roughness caused by a biofilm are available in the literature, nothing similar can be found in the literature for viscous oil coatings.
The objective of the present study, therefore, is twofold. First, a methodology based on computational fluid dynamics (CFD) is tested that would allow for the determination of hydrodynamic roughness under conditions where more conventional methods are difficult to implement or simply unsuitable. For example, velocity profile measurements can be used to infer the hydrodynamic roughness of a surface. In some cases, these cannot be made, usually for economic or technical reasons. In other cases they can be made but, because of flow cell aspect ratio, the measured velocity profile cannot be easily related to the hydrodynamic roughness owing to the presence of strong secondary flows (e.g., Melling and Whitelaw 1976). Then there are cases where the CFD-based methodology can be used instead of standard correlations. For example, the Colebrook formula cannot be used under conditions involving rough-wall asymmetry—such as, two channel walls of different roughness and two smooth walls.
Additionally, hydrodynamic roughness produced by a very viscous coating layer, which is an idealized version of the fouling layer in LPF, is evaluated. The broader motivation, as mentioned previously, is that existing models of LPF do not explicitly address the effect of this fouling layer, even though its contribution to overall friction loss is known to be significant.

Methods

Experimental Facilities and Procedures

A laboratory-scale pipeline flow loop (length 35 m; pipe diameter 25.4 mm), shown schematically in Fig. 1, was used for this study. The primary purpose of the loop was to produce flow through the rectangular test cell, which is shown in Fig. 2.
Fig. 1. Schematic of pipeline flow loop containing the rectangular test cell; overall pipeline length=35  m; pipe diameter=25.4  mm
Fig. 2. Details of the rectangular test cell (dimensions in mm): (a) front view; (b) section A-A
The pipeline loop consisted of a water tank, pump-motor set, flexible connector and damper, heat exchanger, flow cell, filter, and associated copper/flexible tubing. Water from the tank was circulated through the loop by a Moyno 1000-series pump (Moyno, Springfield, Ohio) driven by a variable-frequency drive and 7.5-hp motor (Baldor Industrial Motor, Fort Smith, Arkansas). The pump speed was set to obtain the desired mass flow rate of water. The flexible connector and damper minimized unwanted vibration in the loop. The heat exchanger provided isothermal conditions, with all tests conducted at 20°C. The filter (Arctic P2 filter with 34-μm bag, John Brooks Company, Edmonton, Alberta, Canada) collected oil droplets from the coating layer in the flow cell. A small fraction of the wall-coating oil was stripped from the coating under certain operating conditions, and the filter prevented recirculation of oil droplets through the loop. A Coriolis mass flowmeter (Krohne MFM 4085K Corimass, Type 300G+, KROHNE Messtechnik GmbH, Duisburg, Germany) measured both mass flow rate and temperature.
The rectangular test cell was fabricated from carbon steel and was 2.5 m long and 6 mm thick, with a 25.4×25.4  mm cross section as shown in Fig. 2. The first 1.5 m of this cell served as the entrance length. After the entrance length, there was a flow-visualizing section comprised of two viewing windows. A 9.5-mm-thick stainless steel plate was placed at the bottom of the flow cell. This plate was cut into several segments for the convenience of installation and for conducting surface roughness measurements at the end of each test. The flow cell was equipped with a differential pressure transducer (Validyne P61) for online measurement of pressure loss in the flow cell.
A fully automated Mitutuyo Contracer Contour Measuring System (Model CV-3100H4, Mitutuyo America, Aurora, Illinois) was used to measure the physical roughness of the clean test plate, sandpaper, and oil coating. The Contracer uses a carbide stylus. The weight of this stylus is balanced so that the measuring force (i.e., the effective weight) is only 30 mN. As a result, the surface roughness is not altered or damaged by the stylus. When the stylus travels over a surface, its perpendicular movement (i.e., physical roughness) is quantified. The working principle of the Contracer is analogous to that of a standard contact profilometer. A profilometer usually measures roughness at the micron scale (Flack and Schultz 2010). However, the Contracer can measure roughness over a wider scale, from submicron to millimeter. More importantly, a profilometer is suitable for solid surfaces only whereas a Contracer can also be used for a softer surface if reasonable rigidity is maintained during the measurement. In the present investigation, for example, the oil-coated test plates were in contact with dry ice during the Contracer measurements.
Initially, pressure losses were measured with water flowing through the cell over a clean (unfouled) test plate. Water flow rates were controlled so that the Reynolds number (R) was varied over a range of 104105 (104<R<105). The Coriolis mass flowmeter provided on-line measurements of both mass flow rate and temperature. The total (machine plus precision) error for the two measurements was 0.8 and 9%, respectively (Rushd 2016). Pressure loss measurements were made for different water flow rates to assess the hydrodynamic roughness of the clean walls in the flow cell and then to determine the hydrodynamic roughness of the rough wall (sandpaper, oil coating). Pressure drops were measured under fully developed flow conditions because the entrance length was more than 60Dh, where Dh is the hydraulic diameter. Repeatable steady-state pressure differences measured across this section also indicated fully developed flow. The pressure taps for these measurements were located 800 mm apart over the length of the flow-visualizing section. The first tap was 100 mm downstream of the entrance to this section, and the second one was 100 mm upstream of the exit from the flow cell. The locations of the pressure taps are shown in Fig. 2. The Validyne P61 differential pressure transducer was used to measure the pressure loss. Based on the manufacturer’s specifications and precision measurements made during the experiments, the measurement error was determined to be 5% (Rushd 2016).
Subsequently, flow tests were conducted with sandpaper and a wall coating of heavy oil in the flow cell. For the sandpaper tests, two different sandpapers, of 80 and 120 grit, were glued to the bottom plate in the flow-visualizing section, which was made from a single piece of steel bar as opposed to the set of plates used for the experiments conducted with an oil-coated surface. A slide caliper was used to measure the thickness of the sandpaper. Pressure losses were recorded while water flow rates were varied. Additionally, the topology (surface roughness) of the sandpapers was measured using the Contracer described previously.
For the coating experiments, a viscous heavy oil was used (μo=2.13×104  Pa·s at T=20°C, Husky PG 46-37 300/400A, Husky, Lloydminster, Saskatchewan, Canada). The bottom wall is comprised an assembly of ten plates. Each plate was coated separately with a specific thickness of the viscous oil and placed in the flow cell to form the coating layer of a uniform thickness. The average thickness of the coating layer (tc) was determined by weighing the test plates without and with coating oil. The coated plates were weighed before and after the flow test. The difference between the measured weights was negligible, and therefore tc was considered to be unaffected by the flow rate. The uncertainty associated with the measurement of tc was 11% (Rushd 2016). The two downstream pressure taps (separated by a distance of 450 mm), as shown in Fig. 2, were used for the sandpaper and viscous oil coating tests.
Examples of calculating coating thickness from measured weights of oil on test plates and measuring pressure gradients are presented here for a specific flow condition. The coating thicknesses before and after a flow test are listed in Table 1, and the measured values of pressure gradients (30-s average values) are shown in Fig. 3. These results demonstrate that changes in the coating thickness and corresponding pressure gradients were negligible even after operating at different flow rates over a period of approximately 1.5 h.
Table 1. Example of Coating Thickness (tc=0.5  mm) Determination
Test plate numberWeight of coating oil, mo (g)Coating thickness, tc=mo/ρoAp (mm)Average
Before flow testAfter flow testBefore flow testAfter flow test
11.31.20.50.50.5
21.31.50.50.6
31.31.40.50.5

Note: Oil density, ρo=1,021  kg/m3; area of test plate, Ap=2.54×103  m2.

Fig. 3. Pressure gradients (ΔP/L) measured over time (t) for different mass flow rates of water (mw)
After completing a set of flow tests with a particular wall coating, the coated plates located in the section between the pressure taps were removed from the flow cell. This process took less than 5 min. The test plates were then placed in a freezer maintained at 10°C. At room temperature, the rough surface of the oil coating, which was produced because of the flow past it in the cell, maintained its shape for more than an hour after a test was completed because the oil viscosity was very high. When contour measurements were to be made, a test plate was removed from the freezer and placed on top of a container of dry ice.
The Contracer was then used to measure physical roughness by conducting contour measurements over a large area (xy) of 80×15  mm. The area selected for measurement was located in the center of each test plate (Rushd 2016). The roughness in this area was observed to be unaffected in the course of separating the test plate from the flow cell. The measured area was much larger (1,200  mm2) than the usual test area for measuring roughness on a solid surface, which typically is in the scale of μm2 (Flack and Schultz 2010; Afzal et al. 2013). In addition to assessing the roughness over a large area, the measurements were repeated many times: 29 repeated measurements for test plates coated with a 1-mm layer of oil and 11 measurements for test plates coated with a 0.5-mm layer.

CFD Simulations

The CFD simulations were used to determine hydrodynamic roughness in terms of sand grain equivalent (ks). The unknown hydrodynamic roughness produced on the surface was an input parameter for the simulation. Pressure losses were predicted using a trial-and-error procedure that required iterative specification of the roughness and repeated simulations. The flow conditions in the flow cell were modeled using the CFD software package ANSYS CFX 13.0. A ω-based Reynolds stress model (ω-RSM) was used to model the turbulent flow. The performance of a RSM is generally more accurate, especially in simulating anisotropic flow conditions. Its application for flows in rectangular channels where the geometry induces strong secondary flows was validated by Fletcher et al. (2009) and Amano et al. (2010). Moreover, the superiority of the ω-RSM over the kω model for turbulent flows over rough surfaces (ks1  mm) was demonstrated by Mothe and Sharif (2006). A select number of simulations were also conducted using the k-ω model to show how important selection of the appropriate turbulence model is to the success of this CFD-based method. This comparison is presented in the section “Results and Discussion.”
The most important features of the ω-RSM are described here. The description is based on Fletcher et al. (2009) and the ANSYS CFX-Solver Theory Guide. In this narrative, the differential equations are presented with index notation.
The basic governing equations of turbulent motion for a viscous liquid like water are the Navier-Stokes equations. Turbulent flow fluctuations are included in the model by a time-averaging concept known as Reynolds averaging. In the course of this averaging process, additional terms, known as the Reynolds stresses, appear in the Reynolds-averaged Navier-Stokes (RANS) equations. It is necessary to model the Reynolds stresses for closure of the RANS equations. The RANS equations of continuity and momentum transport for an incompressible Newtonian fluid can be presented in their general forms as follows: for continuity
DρDt=0
(1)
for momentum transport
ρDUiDt=xi(p+23μUkxk)+xj[μ(Uixj+Ujxi)]ρxj(τij)+Si
(2)
where p = static (thermodynamic) pressure; Si = sum of body forces; and τij = fluctuating Reynolds stress contributions.
A number of models are available in ANSYS CFX 13.0 for the Reynolds stresses (τij) in RANS equations. As stated earlier, ω-RSM was selected among the available models as most suitable for this work. In this model, τij is made to satisfy a transport equation. A separate transport equation is solved for each of the six Reynolds stress components of τij. The differential transport equation for Reynolds stress is as follows:
ρDτijDt=ρPijρϕij+23βρωkδij+xk[(μ+μtσk)τijxk]
(3)
The Reynolds stress production tensor Pij is given by
Pij=τikUjxk+τjkUixk,P=12Pkk
(4)
The constitutive relation for the pressure-strain term ϕij in Eq. (3) is expressed as follows:
ϕij=βC1ρω(τij+23kδij)α^(Pij23Pδij)β^(Dij23Pδij)γ^ρk(Sij13Skkδij)
(5)
In this expression, the tensor Dij is given by
Dij=τikUkxj+τjkUkxi
(6)
where the model coefficients are: β=0.09; α^=(8+C2)/11; β^=(8C22)/11; γ^=(60C24)/55; C1=1.8; and C2=0.52.
In addition to the stress equations, the ω-RSM uses the following equations with corresponding coefficients for the turbulent eddy frequency ω and the turbulent kinetic energy k:
ρDωDt=αρωkPkβρω2+xk[(μ+μtσ)ωxk]
(7)
ρDkDt=Pkβρkω+xj[(μ+μtσk)kxj]
(8)
In these equations, Pk is given by
Pk=μt(Uixj+Ujxi)Uixj23Ukxk(3μtUkxk+ρk)
(9)
and the coefficients are
σ*=2;σ=2;β=0.075;α=ββκ2σ(β)0.5=59;κ=0.41;σk=2
In the previously mentioned transport equations, the turbulent viscosity μt is defined as
μt=ρkω
(10)
Usually a wall is treated using the no-slip boundary condition for CFD simulations. Mesh-insensitive automatic near-wall treatment is available for the ω-RSM in ANSYS CFX 13.0. The treatment is meant to control the smooth transition from the viscous sublayer to the turbulent layer through the logarithmic zone. Important features of the near-wall treatment for ω-RSM are outlined in the following paragraphs.
First, in the case of a hydrodynamically smooth wall, the viscous sublayer is connected to the turbulent layer with a log-law region. Velocity profiles for the near-wall regions are as follows: for the viscous sublayer
u+=y+
(11)
and for the log-law region
u+=(1/κ)ln(y+)+BΔB
(12)
where
u+=Ut/uτ;y+=ρΔyuτ/μ=Δyuτ/ν;uτ=(τw/ρ)0.5
In the log-law region, B and ΔB are constants. The value of B is considered as 5.2 and that of ΔB is dependent on the wall roughness. For a smooth wall, ΔB=0. The term Δy, in the definition of y+, is calculated as the distance between the first and second grid points off the wall. Special treatment of y+ in CFX allows one to arbitrarily refine the mesh.
Second, for a hydrodynamically rough wall, the roughness is scaled with the Nikuradse sand grain equivalent (ks). The nondimensional roughness ks+ is defined as ksuτ/ν. A wall is treated as hydrodynamically rough when ks+>70. The value of ΔB is empirically correlated to ks+ as follows:
ΔB=1κln(1+0.3ks+)
(13)
The variable ΔB represents a parallel shift of the logarithmic velocity profile compared with the smooth wall condition.
Third, at the fully rough condition (ks+>70), the viscous sublayer is assumed to be destroyed. The effect of viscosity in the near-wall region is neglected. Finally, the equivalent sand grains are considered to have a blockage effect on the flow. This effect is taken into account by virtually shifting the wall by a distance of 0.5ks.

Simulation Setup

The CFD simulations were conducted for the following flow scenarios: (1) four walls of the flow cell were smooth; and (2) three walls of the cell were smooth and one wall was rough. All computations were performed to obtain steady-state solutions. A typical computational time requirement to obtain a single data point was 45 min using a Lenovo (Morrisville, North Carolina) ThinkStation (processor: Intel Core i5-2500 CPU @3.30 GHz; RAM: 12GB).

Geometry

The geometry of the 3D computational domain was identical to the rectangular test cell. The length (l) of all flow domains tested was 1,000 mm. A length of 2,000 mm was also tested for selected simulations to confirm the length independence of the results. The width (w) was equal to that of the flow cell (25.4 mm). The height (h=15.9tc  mm) was varied depending on the average thickness (tc) of the oil coating on the bottom wall. The tested values of tc were 0.1, 0.2, 0.5, and 1.0 mm. The physical properties of water at 20°C were chosen to match the experiments.

Boundary Conditions

The computational domain had six distinct boundaries where boundary conditions should be prescribed: the inlet, the outlet, and the four walls. At the inlet, the mass flow rate of water and a turbulence intensity of 5% were specified. The same flow rate was also prescribed at the outlet. The no-slip condition was used at the boundaries representing the four walls. These walls were considered to be hydrodynamically smooth for simulating the scenario of the clean flow cell with no wall coating. Flow situations in the cell with a rough surface (wall-coating layer or sandpaper) were simulated by considering the bottom wall as rough and the three other walls as smooth. Specification of ks of the rough wall was required for these simulations. The ks values were unknown for the bottom wall. A trial-and-error procedure was used to determine ks for a particular rough surface. Starting from a low value, the value of ks was increased in increments until the difference between the experimental and simulation results was less than 5%. The final value of ks was considered to be the hydrodynamic roughness of the corresponding rough wall. A more detailed description of the procedure is available in Rushd (2016).

Meshing

The flow geometry was created and meshed with ANSYS ICEM CFD. Since the computational domain was very regular, the software was used to discretize the domain into structured grids. Based on the number of nodes, the meshes tested can be classified as coarse (nodes<50,000), intermediate (50,000<nodes<500,000) and fine (nodes>500,000). The total number of nodes considered to be sufficient for grid independence was 670,200.

Results and Discussion

The CFD-based approach for calculation of hydrodynamic roughness was validated by applying it to analyze data obtained from three independent sets of experiments: clean flow cell tests, sandpaper tests, and tests where surface roughness was produced by a biofouling layer. The experiments involved both solid and soft rough surfaces, which produced equivalent sand grain roughness values in the range of 0–5 mm. Each test case is discussed in detail in the following sections.

Clean Flow Cell Tests

For the flow tests conducted with the clean test cell, the measured pressure gradients with corresponding predictions are presented in Fig. 4 as a function of water velocity, which was calculated from the measured mass flow rate and nominal cross-sectional area of the flow cell. The overall difference between the predicted (using the CFD methodology) and measured values was less than 10% in all cases. Fig. 4 also shows values calculated using a modified form of the Blasius equation, which is suitable for rectangular flow geometries that can induce secondary flows (Jones 1976). All four walls of the flow domain were considered to be hydrodynamically smooth (ks=0) for the simulation. The predicted values of pressure gradients agree quite well with the corresponding measurements for a smooth wall condition. This agreement confirms that the clean walls of the flow cell were smooth during the experiments. These results also help to validate the ability to predict pressure losses in the rectangular flow cell, which may induce secondary flows, using the CFD simulation procedure described earlier.
Fig. 4. Comparison of measured and predicted pressure gradients for a clean flow cell

Sandpaper Tests

The surface topology of the sandpapers obtained from the Contracer measurements was used to determine the equivalent sand grain roughness using a correlation [Eq. (14)] that was proposed by Flack and Schultz (2010) on the basis of two statistical parameters, root mean square roughness (RRMS) and skewness (Rsk)
ks=4.43RRMS(1+Rsk)1.37
(14)
RRMS=i=1nxi2n(xi:datasetofndatapoints)
(15)
Rsk=nn1n2i=1n(xixavg)3[i=1n(xixavg)2]3/2(xavg:arithmeticaverageofxi)
(16)
The RMS values represent the amplitude, and the skewness shows the spatial variation of roughness (King 1980). The values of ks obtained from the Flack and Schultz (2010) correlation, along with the associated statistical parameters obtained from the topology measurements, are provided in Table 2.
Table 2. Hydrodynamic Roughness with Statistical Parameters for the Sandpapers
SandpaperRMS roughness, RRMS (μm)Skewness of roughness, RskSand grain roughness (Flack and Schultz model), ks (mm)
GritThickness (mm)
800.7730.430.5
1200.9550.190.3
For this case study, the estimated values of ks were used as the boundary conditions to predict the measured pressure gradients following the simulation procedure described earlier. The measured values of pressure gradients (ΔP/L)E are compared with the simulation results (ΔP/L)P in Fig. 5, which shows the predictions to vary within ±15% of the measurements. The reasonable agreement shown further validates the CFD methodology introduced here, specifically for a case where the geometry includes asymmetric wall roughness (three smooth walls, one rough wall).
Fig. 5. Comparison of measured and predicted pressure gradients for the sandpaper tests

Biofouling Tests

Andrewartha (2010) conducted experiments using a rectangular flow cell, which was fabricated from plexiglass and had a cross-sectional area of 200×600  mm. Three walls of the custom-built flow cell were smooth, and the fourth wall was coated with a biofouling layer. Two separate parameters (velocity profile and drag force) were measured so that two separate calculations of ks could be made. The velocity profile was measured using both a Pitot tube and laser Doppler velocimetry. The drag force was measured directly with a transducer.
Values of ks, obtained from the velocity profile data (Andrewartha 2010), were then recalculated using the CFD methodology described here. In this example, the measured velocity profiles, rather than the pressure gradients, were predicted using the CFD-based approach. Values of ks were changed iteratively over a series of simulations. The value that could predict the dimensionless velocity (u+) profile within ±10% of the measured values was considered as the representative ks for the corresponding biofouling layer. Fig. 6 is an example of the agreement between the measured and predicted velocity profiles.
Fig. 6. Example of agreement between simulation results and experimental measurements of the velocity profile above a biofouled surface (some data from Sample RP2F5, Andrewartha (2010), was used to calculate results)
The experimentally determined values of ks and the corresponding results of the new CFD-based approach are compared in Table 3. The ks values determined by Andrewartha (2010) from two separate measurements show that significant uncertainty (as high as 100%) can be involved in such determinations. The results obtained from the proposed CFD approach fall within the range of the experimental measurements. This agreement proves that the methodology used in this work is an effective tool for determining ks. It is simple and capable of yielding reliable results.
Table 3. Comparison of the Experimental Hydrodynamic Roughness with Simulation Results for Biofouling Tests
Biofouling sample Andrewartha (2010)Hydrodynamic roughness, ks (mm)
Andrewartha (2010) experimentsCFD simulation (current work)
Method 1: dragMethod 2: velocity profile
RP1F15.735.335.5
RP1F44.473.474.0
RP2F54.372.593.0
SP1F61.030.00 (smooth)0.00 (smooth)

Viscous Oil Coating Tests

For the wall-coating tests, the bottom wall of the flow cell in the flow-visualizing section was coated with viscous oil. The turbulent flow of water changed the topology of the coating surface by producing physical roughness. After an initial period (500  s), visible changes in the roughness were negligible and the pressure losses did not change with time (Fig. 3). The CFD methodology was used to determine the hydrodynamic roughness produced by the wall-coating layers. The results are provided in Table 4. The values of ks were roughly proportional to the values of tc; that is, a thicker coating layer could produce higher roughness under comparable flow conditions. The flow rate was not found to affect the hydrodynamic roughness; a single value of ks could provide reasonable predictions of ΔP/L for different values of mw while tc was constant. This indicates the importance of knowing tc and also indicates a limitation of the CFD methodology developed here: specifically, if some coating is stripped from the wall at higher operating velocities, then ks will also change.
Table 4. Comparison of Measurements with Simulation Results
MeasurementSimulation
Coating thickness, tc (mm)Mass flow rate of water, mw (kg/s)Pressure gradient, ΔP/L (kPa)Pressure gradient, ΔP/L (kPa)Equivalent hydrodynamic roughness, ks (mm)
0.10.62.02.00.1
1.27.67.1
1.815.314.7
0.20.62.22.30.4
1.28.78.3
1.817.617.2
0.50.62.22.81.5
1.210.210.5
1.821.622.2
1.00.63.63.63.5
1.214.013.6
1.828.728.9
To demonstrate the importance of the choice of turbulence model in the performance of the CFD methodology used to determine the hydrodynamic roughness of the viscous oil coating, additional simulations using the k-ω model were conducted. The simulations conducted using the k-ω model were identical to those described earlier, except that, rather than conducting a series of iterative simulations, the ks values from Table 4 were used to predict the pressure gradient. A comparison between the pressure gradient predictions obtained using the CFD methodology and the different turbulence models, for the condition tc=1.0  mm and ks=3.5  mm, is shown in Fig. 7. As expected, the CFD approach, when the k-ω model is incorporated, provides poor predictions of the pressure gradient. In this case, the pressure gradients are under-predicted by approximately 30%; generally, the accuracy would vary widely depending on the extent of surface roughness and on the flow conditions. The limitations of two-equation RANS (e.g., k-ω and k-ϵ) for flows with rough walls and/or anisotropic turbulence are well documented in the literature (e.g., Mothe and Sharif 2006; Fletcher et al. 2009; Amano et al. 2010). The use of the ω-RSM turbulence model, which is a seven-equation model, is particularly important to the success of the CFD methodology described here, considering that the flows all have rough walls. In addition, because of the geometry of the rectangular test cell, strong secondary flows are expected to occur (e.g., Melling and Whitelaw 1976), which are known to be intensified by the presence of a rough wall (Fujita et al. 1989).
Fig. 7. Comparison of simulation results obtained with two different turbulence models, ω-RSM and k-ω (tc=1.0  mm, ks=3.5  mm)
Surface characterization of the clean test plates and the rough wall coatings was conducted with the Contracer. Figs. 8 and 9 show, respectively, the topology of a clean wall and that of a coating layer. Recall that the CFD simulations showed that the clean test plates (no oil coating) behaved as a smooth wall and the coated surface behaved as a rough wall. The difference in the two surfaces can be more fully appreciated by inspecting the 3D plots. The average physical roughness of the clean wall (tc=0) was 2 μm whereas the roughness of the coating surface (tc=1.0  mm) was 266 μm. That is, the oil coating had a physical roughness two orders of magnitude greater than that of the smooth wall.
Fig. 8. Three-dimensional plot of the measured topology of a clean test plate (reprinted from Rushd 2016)
Fig. 9. Three-dimensional plot of the measured topology of a rough wall–coating layer (tc=1.0  mm) (reprinted from Rushd 2016)
It is clear from Fig. 9 that the roughness of the wall-coating layer was 3D in nature. That is, the variation was not consistent in any direction. The most probable reason for this kind of variation in roughness was the presence of secondary flows in the rectangular flow cell. The topological measurements were conducted over a large area (1,200  mm2) to take the 3D nature of roughness into account. Also, the measurements were replicated a number of times (40). Each of these data sets comprised approximately 800×15 data points.
The topological data were used to estimate the values of ks using the correlation [Eq. (14)] proposed by Flack and Schultz (2010). Although the model was developed for solid surfaces, the oil coating tested here had a very high viscosity (μo=2.13×104  Pa·s) and no change in the topology of the coating was observable under steady-state flow conditions. The viscous wall coating did not lose its structure when the flow was stopped or when the test plate was removed from the flow cell. Moreover, the Flack and Schultz (2010) correlation incorporates a term (skewness) to account for the type of nonisometric roughness produced by the flow on the viscous coating.
The equivalent roughness values calculated using the Flack and Schultz model are provided in Table 5. These results show the effect of flow rate on the roughness. The values of ks, which are dependent on the roughness (RRMS) and the skewness (Rsk), tend to decrease with increasing flow rate. Although the variation in the flow rate does not change RRMS appreciably, the increasing flow rate reduces Rsk. In other words, the tested range of Reynolds numbers does not change the magnitude of the roughness significantly, but reduces its spatial variation to some extent. Thus an uncertainty of 30% is associated with the average values of ks assigned to each coating thickness tested in this study. This falls well within the standard range of uncertainty reported in the literature. For example, the uncertainty in determining the hydrodynamic roughness produced by a typical solid surface like steel, glass, or plastic is 20–70% (White 1999; Bhatt 2007), and can be even greater than that if the hydrodynamic roughness is predicted using a friction factor correlation such as the Colebrook correlation. As mentioned earlier, the ks values for different biofouling layers determined using separate experimental measurements (Andrewartha 2010) involved uncertainty in the range of 7–100% (Table 3).
Table 5. Hydrodynamic Roughness and Associated Statistical Parameters for the Flow Cell
Coating thickness, tc (mm)Mass flow rate of water, mw (kg/s)RMS roughness, RRMS (μm)Skewness, RskHydrodynamic roughness, ks (mm)
Surface measurement (Flack and Schultz model)CFD simulation
Flow dependentAverage
0.50.61760.871.81.71.5
1.21810.781.8
1.81810.621.5
1.00.63720.914.03.43.5
1.23150.823.2
1.83160.733.0
It is also clear from the rightmost two columns of Table 5 that good agreement is obtained between the ks determined from the surface measurements and the ks determined from the simulations (reproduced from Table 4) i.e., the values of ks obtained on the basis of two independent methods agree quite well. This agreement supports the CFD approach used in this study to determine the values of ks produced by a viscous layer of wall coating.
One of the main objectives of this study was to implement and test a CFD-based methodology that would allow for the determination of hydrodynamic roughness from bulk measurements—that is, pressure gradient and flow rate. The results presented here for different solid, biofouled, and oil-coated surfaces provide evidence that the CFD approach can be useful, particularly in situations where it is too costly or technically infeasible to make near-wall velocity profile measurements. The other objective was to evaluate the hydrodynamic roughness produced by a viscous oil coating, because this information is not available in the literature and is a critical element of an ongoing effort to develop an improved model of lubricated pipe flow (Rushd and Sanders 2017; M. McKibben, R. Gillies, S. Rushd, and S. Sanders, “A new model to predict pressure losses in the water-assisted pipeline transport of bitumen or heavy oil,” submitted, J. Petrol Explor. Prod. Technol., Springer, Berlin).
In order to gain confidence that the results of this study are comparable with conditions where a pipe wall is coated with viscous oil, some pressure gradient results available in the literature concerning the flow of water through a pipe with a coated wall are analyzed and discussed. Unfortunately, there are very few such data sets available and thus the focus here is on two specific data sets: that of Brauer (1963) and that of Shook et al. (2002). Brauer’s tests were conducted using a 30 mm (diameter) by 600 mm (length) pipe. A relatively thick oil coating (1.5 mm) was used, and then water was pumped through the pipe. One of the shortcomings of these tests is that in many cases oil was stripped from the wall at different operating conditions, meaning that the actual oil coating thickness was often not reported. For one set of data, an equilibrium coating thickness (tc=0.42  mm) was established for a wide range of water flow rates and a relatively constant pressure gradient was measured. These results are shown in Fig. 10, where the pressure measurements have been converted to Fanning friction factors (fF) using
fF=(ΔPL)D2V2ρw
(17)
Fig. 10. Friction factors for water flow through pipes coated internally with viscous oil
The resulting values are plotted against the Reynolds number calculated using the properties of water at 35°C. The coating thickness, the Reynolds number range, and the hydrodynamic roughness reported by Brauer (1963) are provided in Table 6. Also shown in Fig. 10, and described in Table 6, are the results of some tests reported by Shook et al. (2002), where water was pumped through a 53-mm (uncoated pipe diameter) pipeline at a range of flow rates. The pipe was fouled with very viscous bitumen (μo150  Pa·s at 20°C) from the Athabasca oil sands. The coating thickness was known to be approximately 2.5 mm, measured using heat transfer measurements (Schaan et al. 2002). Again, the pressure loss measurements are reported as friction factors in Fig. 10. It can be seen from this figure (and from Table 6) that the Reynolds number range for the flow cell tests discussed here is very similar to that tested by Brauer (1963) and Shook et al. (2002). Also, the friction factors are essentially constant with the Reynolds number, which one would expect for a very rough surface.
Table 6. Comparison of Hydrodynamic Roughness Values for Viscous Coatings: Flow Cell and Pipeline Tests
ResultsRCoating thickness, tc (mm)Hydrodynamic roughness, ks (mm)Ratio, ks/tc
Present study104<R<1050.51.53.0
Present study104<R<1051.03.53.5
Brauer (1963) (Case E)5×104<R<3×1050.421.53.5
Shook et al. (2002)2.5×104<R<1.8×1052.57.02.8
The behavior just described is also consistent with the results of this study; that is, coating thickness produces a relatively constant value of ks over a range of conditions. Table 6 summarizes the ratio of hydrodynamic roughness (ks) to coating thickness (tc) for the present study and for the tests of Brauer (1963) and Shook et al. (2002). It can be seen that the ratio (3) is very similar for the rectangular flow cell and for the pipe flow tests, providing confidence that the measurements made in the flow cell reasonably represent pipe flow, at least for the range of conditions tested. In fact, the application of the CFD methodology to pipe flows with viscous coatings at much higher Reynolds numbers than those studied here, which was the subject of a separate investigation (Rushd and Sanders 2017), has been shown to provide excellent results and represents another step in the effort to produce an accurate, mechanistic model of LPF.

Summary and Conclusion

In this study, a CFD-based approach was used to determine the hydrodynamic roughness of a rough surface based on bulk measurements—flow rate and frictional pressure loss. Typically, velocity profiles must be measured above the rough surface. This approach is known to be time-consuming and, for a more complicated flow, such as lubricated pipe flow, would be technically infeasible. Another alternative is the use of friction factor correlations, such as the Colebrook-White equation, but this approach can provide very large changes in the value of ks for relatively small changes in the measured pressure loss. Therefore, the CFD-based approach described here is viewed as a viable alternative. Compared with the approach where velocity profiles are measured to determine ks, the CFD approach is extremely time- and cost-effective. The CFD approach was tested against the velocity profile method applied to determine the hydrodynamic roughness of a biofouled surface. Good agreement was obtained. Subsequently, the CFD methodology was applied to determine the hydrodynamic roughness of two different sandpapers. Here, the CFD results were compared with those obtained using the correlation of Flack and Schultz (2010), which uses the statistics of the surface topology. Again, good agreement was observed between the two methods. The performance of the CFD methodology is highly dependent on the selection of a turbulence model. Here, the 7-equation ω-RSM model, which was developed for anisotropic flow conditions and/or flow over rough surfaces, was shown to be superior to the two-equation k-ω model for the range of conditions studied. Although the CFD methodology provided reasonable predictions here, other turbulence models should be tested, particularly for very high Reynolds number conditions.
The other primary focus of this study was to gain new understanding of the hydrodynamic behavior of a viscous oil coating, given that such coatings are a dominant mechanism for pressure loss in LPF. The present study was intended to provide some preliminary information about the relationships among coating thickness, flow Reynolds number, and hydrodynamic roughness. Additionally, the surface topology of the viscous oil coatings was measured. In the rectangular flow cell tested here, it was discovered that the hydrodynamic roughness was nearly independent of R (for 104<R<105) and approximately proportional to the coating thickness. Similar findings were observed when the coated-wall pipe flow data of Brauer (1963) and Shook et al. (2002) were reanalyzed, giving confidence that the rectangular flow cell measurements made here are consistent with pipeline flow behavior. One important limitation is that, if the oil coating is partially stripped from the wall, these results will not hold. Therefore, additional tests must be done where oils with different viscosities are tested. Additionally, larger pipes (i.e., higher Reynolds numbers) must be tested. Also, before the results of this study can be applied with confidence to LPF applications, the impact of oil droplet interaction between the mean flow and the coated wall should be investigated. This could be accomplished by conducting experiments similar to those described here, except that the flowing fluid should consist of oil droplets suspended in water.
Finally, this is the first time, to the authors’ knowledge, that the Flack and Schultz (2010) correlation has been applied to viscous oil coatings. The surface topology measurements made here, and the use of the Flack and Schultz correlation to convert those measurements to a value of hydrodynamic roughness, proved to be very valuable. It was determined that, for a given coating thickness, the amplitude of the surface roughness did not change substantially with Reynolds number; however, as the Reynolds number increased, the roughness became more structured. Again, it would be most useful to determine in the future how oil viscosity affects this relationship.

Notation

The following symbols are used in this paper:
A
cross-sectional area perpendicular to flow direction (m2);
Ap
surface area on test plate (m2);
D
diameter (m);
Dh
hydraulic diameter (m);
Dij
a tensor;
h
height of flow domain (m);
k
turbulent kinetic energy (J/kg);
ks
Nikuradse sand grain equivalent roughness (m);
l
length of flow domain (mm);
mw
mass flow rate of water (kg/s);
Pij
Reynolds stress production tensor;
p
static (thermodynamic) pressure;
R
Reynolds number;
Ra
average roughness (m);
RRMS
root mean square roughness (m);
Rsk
skewness of roughness;
Rz
peak-to-valley average roughness (m);
Si
sum of body forces (N);
T
temperature (°C);
tc
average thickness of wall-coating layer (m);
Ui
velocity vector (m/s);
uτ
friction/shear velocity (m/s);
u+
dimensionless velocity;
V
average velocity (m/s);
W
width of flow domain (m);
y+
dimensionless distance from wall;
δij
identity matrix or Kronecker delta function;
ΔP
pressure drop (Pa);
ΔP/L
pressure gradient (Pa/m);
μ
viscosity (Pa·s);
μo
viscosity of coating oil (Pa·s);
μt
turbulent eddy viscosity (Pa·s);
ρ
density (kg/m3);
ρo
density of oil (kg/m3);
τij
stress tensor;
ϕij
pressure-strain tensor; and
ω
turbulent eddy frequency (1/s).

Acknowledgments

This research was conducted with support from the NSERC Industrial Research Chair in Pipeline Transport Processes (RSS). The authors acknowledge the support of Canada’s Natural Sciences and Engineering Research Council (NSERC) and the sponsoring companies: Canadian Natural Resources Ltd., CNOOC-Nexen Inc., Saskatchewan Research Council’s Pipe Flow Technology Centre, Shell Canada Energy, Suncor Energy, Syncrude Canada Ltd., Total, Teck Resources Ltd., and Paterson & Cooke Consulting Engineers Ltd.

References

Afzal, N., Seena, A., and Bushra, A. (2013). “Turbulent flow in a machine honed rough pipe for large Reynolds numbers: General roughness scaling laws.” J. Hydro-environ. Res., 7(1), 81–90.
Amano, R. S., Guntur, K., and Lucci, J. M. (2010). “Computational study of gas turbine blade cooling channel.” Proc., 14th Int. Heat Transfer Conf., Vol. 5, ASME, New York.
Andrewartha, J. M. (2010). “The effect of freshwater biofilms on turbulent boundary layers and the implications for hydropower canals.” Ph.D. thesis, Univ. of Tasmania, Hobart, Australia.
ANSYS version 13.0 [Computer software]. ANSYS, Canonsburg, PA.
Arney, M. S., Bai, R., Guevara, E., Joseph, D. D., and Liu, K. (1993). “Friction factor and holdup studies for lubricated pipelining. I: Experiments and correlations.” Int. J. Multiphase Flow, 19(6), 1061–1076.
Arney, M. S., Ribeiro, G. S., Guevara, E., Bai, R., and Joseph, D. D. (1996). “Cement-lined pipes for water lubricated transport of heavy oil.” Int. J. Multiphase Flow, 22(2), 207–221.
Barton, A. F., Sargison, J. E., Walker, G. J., Osborn, J. E., and Brandner, P. A. (2005). “A baseline study of the effect of freshwater biofilms in hydraulic conduits.” Proc., XXXI IAHR Congress, International Association for Hydro-Environment Engineering and Research, Beijing, 992–1002.
Bayazit, M. (2010). “Free surface flow in a channel of large relative roughness.” J. Hydraul. Res., 14(2), 115–126.
Bhatt, C. P. (2007). “Assessment of uncertainty in equivalent sand grain roughness methods.” M.Sc. thesis, Univ. of Alabama, Birmingham, AL.
Brauer, H. (1963). “Flow resistance in pipes with ripple roughness.” Chemische Zeitung (Chem. Rev. Eng.), 87(6), 199–210.
CAPP (Canadian Association of Petroleum Production). (2015). “Crude oil: Forecast, market and transportation.” ⟨http://www.capp.ca⟩ (Dec. 15, 2015).
Crivelaro, K. C. O., Damacena, Y. T., Andrade, T. H. F., Lima, A. G. B., and Farias Neto, S. R. (2009). “Numerical simulation of heavy oil flows in pipes using the core-annular flow technique.” WIT Trans. Eng. Sci., 63, 193–203.
Flack, K., and Schultz, M. P. (2010). “Roughness scaling parameters in the fully-rough regime.” Proc., IUTAM Symp. on the Physics of Wall-Bounded Turbulent Flows on Rough Walls, Springer, Dordrecht, Netherlands, 99–104.
Fletcher, D. F., Geyer, P. E., and Haynes, B. S. (2009). “Assessment of the SST and omega-based Reynolds stress models for the prediction of flow and heat transfer in a square-section u-bend.” Computat. Therm. Sci., 1(4), 385–403.
Fujita, H., Yokosawa, H., and Hirota, M. (1989). “Secondary flow of the second kind in rectangular ducts with one rough wall.” Exp. Therm. Fluid Sci., 2(1), 72–80.
Govindarajan, R., and Sahu, K. C. (2014). “Instabilities in viscosity-stratified flow.” Ann. Rev. Fluid Mech., 46, 331–353.
Guevara, E., Gonzal, J., and Nunez, G. (1998). “Highly viscous oil transportation methods in the Venezuela oil industry.” Proc., 15th World Petroleum Congress, World Petroleum Council, London, 495–502.
Hart, A. (2014). “A review of technologies for transporting heavy crude oil and bitumen via pipelines.” J. Petrol Explor. Prod. Technol., 4(3), 327–336.
Ho, W. S., and Li, N. N. (1994). “Core-annular flow of liquid membrane emulsion.” AICHE J., 40(12), 1961–1968.
Hooper, A. P., and Boyd, W. G. C. (1987). “Shear-flow instability due to a wall and a viscosity discontinuity at the interface.” J. Fluid Mech., 179(1), 201–225.
Jean, F. G., Sébastien, G., Guy, M., and Armelle, S. (2005). “Prospective analysis of the potential non-conventional world oil supply: Tar sands, oil shales and non-conventional liquid fuels from coal and gas.”, European Commission, Seville, Spain.
Jones, O. C. (1976). “An improvement in the calculation of turbulent friction in rectangular ducts.” J. Fluids Eng., 98(2), 173–180.
Joseph, D. D., Bai, R., Mata, C., Sury, K., and Grant, C. (1999). “Self-lubricated transport of bitumen froth.” J. Fluid Mech., 386, 127–148.
Kandlikar, S. G., Schmitt, D., Carrano, A. L., and Taylor, J. B. (2005). “Characterization of surface roughness effects on pressure drop in single-phase flow in minichannels.” Phys. Fluids, 17(10), 100606.
King, T. G. (1980). “RMS, skew and kurtosis of surface profile height distributions: Some aspects of sample variation.” Precis. Eng., 2(4), 207–215.
Kushnir, R., Segal, V., Ullmann, A., and Brauner, N. (2014). “Inclined two-layered stratified channel flows: Long wave stability analysis of multiple solution regions.” Int. J. Multiphase Flow, 62, 17–29.
Lambert, M. F., Edwards, R. W. J., Howie, S. J., Gilio, B. B. D., and Quinn, S. P. (2009). “The impact of biofilm development on pipe roughness and velocity profile.” Proc., World Environment Water Resources Congress 2009, ASCE, Reston, VA, 122–134.
McKibben, M. J., Gillies, R. G., and Shook, C. A. (2000). “Predicting pressure gradients in heavy oil-water pipelines.” Can. J. Chem. Eng., 78(4), 752–756.
Melling, A., and Whitelaw, J. H. (1976). “Turbulent flow in a rectangular duct.” J. Fluid Mech., 78(2), 289–315.
Mothe, K. K., and Sharif, M. A. R. (2006). “Assessment of the omega based models in the prediction of incompressible turbulent flows over flat rough surfaces.” Collection of Technical Papers—44th AIAA Aerospace Sciences Meeting, American Institute of Aeronautics and Astronautics, Reston, VA, 8707–8715.
Nunez, G. A., Rivas, H. J., and Joseph, D. (1998). “Drive to produce heavy crude prompts variety of transportation methods.” Oil Gas J., 96(43), 59–67.
Picologlou, B. F., Characklis, W. G., and Zelver, N. (1980). “Biofilm growth and hydraulic performance.” J. Hydraul. Div., 106(5), 733–746.
Rodriguez, O. M. H., Bannwart, A. C., and de Carvalho, C. H. M. (2009). “Pressure loss in core-annular flow: Modeling, experimental investigation and full-scale experiments.” J. Petrol. Sci. Eng., 65(1–2), 67–75.
Rushd, S. (2016). “A new approach to model friction losses in the water-assisted pipeline transportation of heavy oil and bitumen.” Ph.D. thesis, Univ. of Alberta, Edmonton, Canada.
Rushd, S., and Sanders, R. S. (2017). “A parametric study of the hydrodynamic roughness produced by a wall-coating layer of heavy oil.” J. Petrol. Sci., 14(1), 155–166.
Sanders, R. S., Ko, T., Bai, R., and Joseph, D. D. (2004). “Factors governing friction losses in self-lubricated transport of bitumen froth. I: water release.” Can. J. Chem. Eng., 82(4), 735–742.
Saniere, A., Henaut, I., and Argillier, J. F. (2004). “Pipeline transportation of heavy oils, a strategic, economic and technological challenge.” Oil Gas Sci. Technol., 59(5), 455–466.
Schaan, J., Sanders, R. S., Litzenberger, C., Gillies, R. G., and Shook, C. A. (2002). “Measurement of heat transfer coefficients in pipeline flow of Athabasca Bitumen froth.” Proc., 3rd North American Conf. on Multiphase Technology, BHR Group, Banff, Canada, 25–38.
Schultz, M. P. (2007). “Effects of coating roughness and biofouling on ship resistance and powering.” Biofouling J. Bioadhesion Biofilm Res., 23(5), 331–341.
Shook, C. A., Gillies, R. G., and Sanders, R. S. (2002). Pipeline hydrotransport with applications in the oil sand industry, Saskatchewan Research Council Pipe Flow Technology Centre, Saskatoon, Canada.
Tilley, B. S., Davis, S. H., and Bankoff, S. G. (1994). “Linear stability of two-layer fluid flow in an inclined channel.” Phys. Fluids, 6(12), 3906–3922.
White, F. M. (1999). Fluid mechanics, 4th Ed., McGraw-Hill, Boston, 349.

Information & Authors

Information

Published In

Go to Journal of Hydraulic Engineering
Journal of Hydraulic Engineering
Volume 144Issue 2February 2018

History

Received: Jul 27, 2016
Accepted: May 15, 2017
Published online: Dec 15, 2017
Published in print: Feb 1, 2018
Discussion open until: May 15, 2018

Authors

Affiliations

Sayeed Rushd
Ph.D. Student, Dept. of Chemical and Materials Engineering, Univ. of Alberta, Edmonton, AB, Canada T6G 2V4.
Ashraful Islam
Research Associate, Dept. of Chemical and Materials Engineering, Univ. of Alberta, Edmonton, AB, Canada T6G 2V4.
Professor, Dept. of Chemical and Materials Engineering, Univ. of Alberta, Edmonton, AB, Canada T6G 2V4 (corresponding author). ORCID: https://orcid.org/0000-0002-5611-0423. 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