Sie sind auf Seite 1von 45

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/329448323

Pressure loss and friction factor in non-Newtonian mine paste backfill:


Modelling, loop test and mine field data

Article  in  Powder Technology · December 2018


DOI: 10.1016/j.powtec.2018.12.029

CITATIONS READS

0 89

6 authors, including:

mc Kermani Ferri Hassani


McGill University McGill University
8 PUBLICATIONS   13 CITATIONS    141 PUBLICATIONS   776 CITATIONS   

SEE PROFILE SEE PROFILE

A.P. Sasmito
McGill University
121 PUBLICATIONS   1,246 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Numerical investigation of impinging drying with intermittent input View project

Heat, mass and momentum transfer View project

All content following this page was uploaded by A.P. Sasmito on 10 April 2019.

The user has requested enhancement of the downloaded file.


Pressure loss and friction factor in non-Newtonian mine paste backfill:
modelling, loop test and mine field data.

Bhargav Bharathan a, Maureen McGuinness a, Sharun Kuhar b, Mehrdad


Kermani, Ferri P. Hassani a and Agus P. Sasmito a,*

a
Department of Mining and Materials Engineering Department, McGill University, 3450
University Street, Montreal, QC Canada H3A 0E8

b
Mechanical Engineering Department, Indian Institute of Technology, Kanpur, India

*Corresponding author: Agus P. Sasmito

E-mail address: agus.sasmito@mcgill.ca

1
Pressure loss and friction factor in non-Newtonian mine paste backfill:
modelling, loop test and mine field data

Abstract
The prediction of pressure loss along paste backfill pipeline system requires

careful consideration to determine the flowability and pumping requirement. Several

friction factor correlations are available in literature to predict pressure loss for paste;

however, none of them was specifically developed and validated for mine paste backfill.

This study addresses the selection of friction factor correlations in hydraulic model to be

tested and validated against pressure loss from laboratory scale pipe loop measurement,

in-situ mine field data and computational fluid dynamics model. Statistical analysis was

performed to quantify the errors. The results suggest that Swamee and Aggarwal friction

factor correlation can accurately predict the pressure loss along the paste backfill pipeline

in the mine field within 5% errors.

Keywords: paste backfill; friction factor; rheology; computational fluid dynamics; non-
Newtonian; Bingham.

1. Introduction
Backfilling is a mine waste management method that sends the tailings back

underground to provide stability to the mine walls for successful extraction from adjacent

stopes [1]. The open stope voids are filled with backfill to provide stability for the

adjacent work areas and reducing the risk of local and regional ground failure from

collapse and caving in of the structure. Backfill is typically made from waste rock or

dewatered tailing residues and is often mixed with cement to achieve moderate strengths.

It is delivered to stopes via trucks or by pumping a dense tailings slurry or paste through

boreholes and pipelines. The backfill is made by mixing tailings (sometimes blended with

sand or other aggregate), water and a binder to produce the desirable structural properties

2
[1][2]. The fine tailings are mixed with small amounts of water to enable transport by

pipeline without producing a settling slurry. The resulting fluid is a thick paste with high

mass concentration (usually greater than 70 weight % solids), with a minimum fines

content of at least 20% <20 µm and exhibit non-Newtonian fluid characteristics

[3][4][5][6][7]. The amount of water added is controlled to provide a balance between the

backfill fluidity and its mechanical properties which is one of its key performance

properties [8].

Backfill is transported from the processing plant on the surface into the mine via

pipelines under gravitational force or pumping pressure [9]. The pressure variation along

the pipe length is important in the pipeline design when evaluating the need for pumps

and booster stations in a pipeline and in the selection of the optimum pipe size. A key

parameter in the pressure loss is the friction factor which is commonly expressed by the

Fanning friction factor or the Darcy-Weisbach friction factor that differ from each other

by a factor of 4 as shown in Eqs. (12) and (13). Friction factor correlations for dilute and

moderately dense slurry flows for Newtonian and non-Newtonian fluids have been

extensively studied in past literature while studies involving high concentration slurry

flows are relatively recent. It is the friction factor for high concentration, non-Newtonian

fluids that is of interest for paste backfill.

Senapati and Mishra simulated 60 to 70% by weight fly ash slurry using the non-

Newtonian power law model and define the friction loss using the fanning friction factor

[10]. Liu et al studied the rheology of 50 to 70% by weight coal sludge slurry [11]. They

used the Herschel-Bulkley model to study the rheology of their flow but make no mention

of friction factor. Assefa and Kaushal did a comparative study of several friction factor

correlations for smooth pipes for 60 to 70% by weight slurry using experimental data

[12][13][14]. They found Wilson-Thomas’ [15] and Morrison’s [16] correlations to be

3
the best in agreement with experimental data for the Reynolds number range 200 to

40,000. Wu simulated 70 to 80% by weight coal gangue-fly ash slurry and compared it

with loop test experiments [4][17]. In their study the authors employed Swamee &

Aggarwal’s friction factor model [18]. Despite there being a few studies tackling high

concentration slurry flows, none of them focus on the friction factors at low Reynolds

numbers. No comparison between experiments, friction factor correlations and numerical

simulations were found together for Bingham plastic fluids. The only friction factor study

which compared some of these elements was done by Assefa and Kaushal, but the study

doesn’t focus on high concentrations at low Reynolds numbers, which is very typical of

paste backfill flows.

Qi et al. experimentally observed the pressure drop in pipe flow of cemented paste

backfill in complex circuits and developed a prediction model using gradient boosting

regression tree (GBRT) [19]. The authors achieved a 0.999 and 0.998 correlation

coefficient and declared the circuit shape as the most sensitive variable followed by solid

fraction, inlet velocity and cement-tailings ratio.

This study aims to make a comparison between available friction factor

correlations in the low Reynolds numbers regime for non-Newtonian flows. They are in

turn compared with experimental data and numerical results with the objective of finding

the best suited friction factor relationship for this flow regime. The experimental data are

obtained from a laboratory-scale flow loop experiment using tailings and mine process

water from two mines in Canada, henceforth referred to as Mine-A and Mine-B. The

numerical study is performed using CFD by adopting the two-parameter Bingham plastic

model to represent paste backfill rheology. A comparison of generalized Reynolds

number as proposed by K. Madlener, B. Frey and H.K. Ciezki [20] against the Newtonian

4
Reynolds number that is conventionally used for non-Newtonian fluids is done to assess

its compatibility in paste backfill applications.

2. Paste Backfill Rheology


The prediction of the flow characteristics is based on accurate modelling of the

rheology of the backfill. For the fluid model, the rheological parameters are determined

by curve fitting the loop test pseudo shear results with a non-Newtonian fluid equation.

There are several Non-Newtonian fluid models, but the Herschel-Bulkley model and it’s

subgroup, the Bingham plastic model, are found to best represent the nature of high

concentration slurries such as paste backfill [21]. Though it has been observed that in

some cases the three parameter Herschel-Bulkley model better fits the fluid rheogram

[6][22], the Bingham plastic model is more popular because of its simplicity

[4][5][17][23] and it is used in this study to determine the rheology of paste backfill.

The composition of the tailings from Mine-A and Mine-B are detailed below in

Table 1. The particle size distribution of tailings is shown in Fig.1 and Table 2. Mine-A

tailings has a wider range of particle sizes than Mine-B. Mine B tailings are very fine

with 80% of the material passing 20 µm. Particle sizes have significant impact on the

pressure drop across a pipe and is dependent on solid fraction and flow rate. Finer

particles have low sediment velocity and can be easily kept in circulation in comparison

to larger particles [24].

3. Flow Loop Test


The flow loop test experiment is employed to measure the pressure gradient along

the length of the pipe and characterize the fluid flow behaviour in pipes. A schematic

arrangement of the setup is shown below in Fig. 2. The system is in total 12 m long with

pipes running 8.2 m and the rest comprising of the positive displacement pump and

5
hopper. The pipe inner diameter is 0.045 m (1.83 in). The temperature of the backfill in

the pipe is controlled via a heat exchanger which is connected to a temperature control

mechanism. The mine tailings and water are homogeneously mixed in the hopper and

introduced into the system. The positive displacement pump circulates the fluid using a

variable frequency drive (VFD) to control the mass flow rate of the fluid.

Note: PD – Positive displacement pump; HP – Hopper; P1, P2, P3 and P4 –

Pressure gauges; HE – Heat Exchanger; FM – Flow meter.

There are 4 pressure gauges installed; 2 upstream and 2 downstream. The pipe

length L between gauges P1-P2 and P3-P4 is 1500 mm. From the data collected by the

pressure gauges, the pressure drop due to the resistance in flow is obtained by removing

the gravitational head loss as detailed below in Eqs. (1) and (2).

∆𝑃12 = (𝑃1 − 𝜌𝑔𝐿) − 𝑃2 (1)

∆𝑃34 = 𝑃3 − (𝑃4 − 𝜌𝑔𝐿) (2)

Applying Eqs. (1) and (2) to the loop test pressure readings, it was observed that

ΔP12 was equal to ΔP34 within an error of 0.1%. A single value of pressure gradient is

obtained by averaging ΔP12 and ΔP34 as shown below in Eq. (3). Prior to collecting the

pressure data, the backfill is permitted to circulate for sufficient time till a steady-state

flow is attained in the pipeline.

∆𝑃 1 ∆𝑃12 ∆𝑃34
= ( + ) (3)
𝐿 2 𝐿 𝐿

The backfill is first mixed at the thickest desired concentration and brought to the

desired temperature using the heat exchanger. It is pumped at various speeds using the

VFD during which time the pressures in the loop are recorded along with the flowrate.

The paste is diluted by adding processed mine water and the test is repeated for lower

solid concentrations. At each solid concentration, a sample is weighed and dried in an

oven to determine the exact solid fraction of the backfill used in each trial. Before and

6
after the completion of each test, water is flushed through the pipes to clean out residues

and settlements from previous trials. This entire process is repeated at three different

temperatures, 16°C, 26°C and 35°C for both mine tailings.

For determining the rheological properties of the paste backfill, the Buckingham-

Reiner equation shown below in Eq. (4) is used. The shear rate 𝛾 ′ from this is equated

with the pseudo shear rate in Eq. (5). The wall shear stress expressed in terms of pressure

loss is displayed in Eq. (6). The values from Eqs. (5) and (6) are curve-fit into Eq. (7)

using linear regression to determine the Bingham plastic viscosity 𝜇𝑃 and the Bingham

yield stress 𝜏𝐵 .

𝜏𝑊 4 𝜏𝐵 1 𝜏𝐵 4
𝛾′ = [1 − ( ) + ( ) ] (4)
𝜇𝑃 3 𝜏𝑊 3 𝜏𝑊

8𝑉
𝛾′ = (5)
𝐷
𝐷 ∆𝑃
𝜏𝑊 = (6)
4 𝐿

4. Reynolds Number
The dimensionless Reynolds number is the ratio of inertial forces to viscous forces

experienced by a fluid while flowing. In practice it is used to identify the flow regime as

laminar, transitional or turbulent. If the inertial forces, which resist a change in velocity

is dominant then the flow is turbulent and if the viscous forces, which provide a resistance

to the flow is dominant, then the flow is laminar. In an internal flow through a pipe of

hydraulic diameter 𝐷, the Newtonian Reynolds number is written as shown below in Eq.

(7).

𝜌𝑢𝐷
𝑅𝑒𝑛𝑒𝑤𝑡 = (7)
𝜇

7
However, this relies on the assumption that the dynamic viscosity 𝜇 is constant.

This is not the case in non-Newtonian fluids. In 1955, Metzner and Reed developed the

generalized Reynolds number for non-Newtonian fluids.

𝜌𝑢2−𝑛 𝐷𝑛
𝑅𝑒𝑔𝑒𝑛 = (8)
8𝑛−1 𝜇𝑝

For Newtonian fluids, 𝜇𝑝 = 𝜇 and 𝑛=1 and Eq. (8) reduces to Eq. (7). For Bingham

plastic fluids, the Eq. (8) reduces to the Bingham Reynolds number.

𝜌𝑢𝐷
𝑅𝑒𝐵 = (9)
𝜇𝑝

In 2009, K. Madlener, B. Frey and H.K. Ciezki derived an extended version of

the generalized Reynolds number for Herschel-Bulkley type fluids. The authors

compared the results with paraffin-based fluids and noted good agreement between

experimental data and theoretical Darcy friction factor for 𝑅𝑒𝑔𝑒𝑛 𝐵 =1000. At higher

Reynolds numbers lower theoretical friction factor values were observed. The generalized

Bingham Reynolds number is shown below in Eq. (10).

𝜌𝑢𝐷
𝑅𝑒𝐺𝐸𝑁 𝐵𝐼𝑁𝐺𝐻𝐴𝑀 = (10)
𝜏 𝐷 (3𝑚 + 1)
( 8𝐵 ) ( 𝑢 ) + 𝜇𝑝
(4𝑚)

8𝑢
𝜇𝑝 ( 𝐷 )
𝑚= (11)
8𝑢
𝜏 𝐵 + 𝜇𝑝 ( 𝐷 )

Like the Reynolds number, the Hedstrom number is a measure of the regime

characterizing a flow. It is a dimensionless value that combines the yield stress to viscous

stress relationship of Bingham number with the flow properties of Reynolds number. At

𝐻𝑒 = 0 the fluid is Newtonian and it increases proportionately with Bingham yield stress.

A large Hedstrom number shows a strong non-Newtonian nature.

𝜌𝑢𝐷 𝜏𝐵 𝐷 𝜌𝐷2 𝜏𝐵
𝐻𝑒 = 𝑅𝑒𝐵 × 𝐵 = [ ]×[ ]= (12)
𝜇𝑝 𝜇𝑝 𝑢 𝜇𝑝 2

8
It is typically used to calculate a flow’s critical Reynolds number to represent the

point at which the change from laminar to turbulent flow occurs. In the work by Swamee

and Aggarwal [18], the authors arrive at a relationship between critical Reynolds number

and Hedstrom number for the range, 1 ≤ 𝐻𝑒 ≤ 108 , shown below in Eq. (13).

𝐻𝑒 0.35
𝐻𝑒 = 2100 [1 + ] (13)
3600

5. Friction Factor
The concept of friction factor originated from the phenomenological Darcy-

Weisbach equation for pipe flow:

𝛥𝑃 2𝐷
𝑓= (14)
𝐿 𝜌𝑢2

It is not to be confused with Fanning’s friction factor, which is one-fourth of

Darcy’s friction factor.

𝛥𝑃 𝐷
𝑓= (15)
𝐿 2𝜌𝑢2

The first non-Newtonian friction factor study in the laminar regime was

completed by Metzner and Reed for different fluids with their rheology modelled by the

power law [25]. A study performed by Garcia and Steffe [26], compared friction factors

of different kinds of non-Newtonian fluids, including Bingham fluids, and concluded that

the results from friction factor models available at that time differed significantly. As the

study focused on higher Reynolds numbers and lacked experimental backing it is not

directly applicable to the present study. The most recent comparative friction factor study

was carried out by Assefa and Kaushal [12]. They reviewed many of the available friction

factor models for non-Newtonian fluids including Hagen-Poiseuille [27], Blasius [28],

Buckingham-Reiner [29], Colebrook [30], Prandtl [31], Darby and Melson [32], Slatter

9
[33], Danish & Kumar [34], Swamee and Aggarwal [18], Wilson and Thomas [15] and

Morrison [16]. But, the study doesn’t focus on high concentration slurries at low flow

rates that represent paste backfill flow. Consequently, the present study was initiated to

compare the relevance of friction factor models for Bingham fluids at high concentration

and low flow rates, using the Assefa and Kaushal study as a starting point for statistical

evaluation and selection of potential models.

The friction factor models discussed in this study are those put forth by Hagen-

Poiseuille [27], Swamee and Aggarwal [18], Danish & Kumar [34], J.S. Curtis [35],

Darby and Melson [32], Ihle and Tamburrino [36] and Hanks [37]. The friction factor

models are discussed in detail below starting with the Buckingham-Reiner equation.

5.1. Buckingham-Reiner
An exact description of Darcy-Weisbach friction loss for Bingham plastics for a

fully developed flow was first published by E. Buckingham [29]. The Eq. (16) below is

written in a dimensionless form using the Bingham Reynolds and Hedstrom numbers.

Solving this equation for an exact analytical solution is complex due to the fourth order

polynomial equation in 𝑓. Several researchers have attempted to develop explicit

approximations of this equation.

64 1 𝐻𝑒 64 𝐻𝑒 4
𝑓= [1 + − ( 3 7 )] (16)
𝑅𝑒𝐵 6 𝑅𝑒𝐵 𝑅𝑒𝐵 𝑓 𝑅𝑒𝐵

5.2. Hagen-Poissuille
This is the simplest and most common formula used to determine friction factor

in a laminar flow. It assumes an incompressible Newtonian fluid that is non-accelerating.

The formulation is simple and fits accurately for the laminar flow of Newtonian fluids.

10
The model ignores the Bingham yield stress and relies solely on the Bingham plastic

viscosity.

64
𝑓= (17)
𝑅𝑒𝐵

5.3. Swamee and Aggarwal


This correlation is used to explicitly calculate the friction factor for Bingham

plastic fluids flowing in pipes in the laminar flow regime. It is an approximation of the

Buckingham-Reiner equation that covers a wide range of Re and He values aimed at

avoiding the use of complex and time-consuming iterative solutions. The model is a good

solution to the principal Buckingham equation and depending on how closely a fluid

shows Bingham behaviour, this model can be used. An exact Bingham fluid would be

exceedingly well governed by this model but the discrepancy in our case, though less

compared to other models, is due to the approximation that the backfill is modelled as a

Bingham fluid.

𝐻𝑒 1.143
64 10.67 + 0.1414 (𝑅𝑒 )
𝐵
𝑓= + 1.16
(18)
𝑅𝑒𝐵 𝐻𝑒
(1 + 0.0149 (𝑅𝑒 ) ) 𝑅𝑒𝐵
𝐵

5.4. Danish and Kumar


The Danish and Kumar model provides an explicit procedure to calculate friction

factor for the flow of Bingham plastic fluids in smooth pipes by applying Adomian

Decomposition Method (ADM) and the more effective Restarted ADM (RADM) on the

Colebrooke Eq. [30] for turbulent flows and on the Buckingham Eq. [29] for laminar

flows. The explicit relations obtained from RADM were highly accurate for turbulent

flows (within 0.005% error) and fairly accurate for laminar flows (within 5.2% error). In

11
the laminar regime, errors were reported at high Reynolds (Re=1x105) and Hedstrom

numbers (He=1x1010).

4𝐾2
𝐾1 + 3
𝐾𝐾
(𝐾1 + 4 1 2 )
𝐾1 + 3𝐾2
𝑓= (19)
3𝐾2
1+ 4
𝐾𝐾
(𝐾1 + 4 1 2 )
𝐾1 + 3𝐾2

16 16𝐻𝑒
𝐾1 = + (20)
𝑅𝑒𝐵 6𝑅𝑒𝐵 2

16𝐻𝑒 4
𝐾2 = − (21)
3𝑅𝑒𝐵 8

5.5. Jennifer Sinclair Curtis


In the book Introduction to Particle Technology, J.S. Curtis arrives at an

expression for mean velocity of a Bingham plastic fluid in terms of yield stress and wall

shear stress.

𝐷𝜏𝐵 4 𝜏𝐵 1 𝜏𝐵 4
𝑢𝑚𝑒𝑎𝑛 = [1 − + ( ) ] (22)
2𝜇𝑝 3 𝜏𝑤 3 𝜏𝑤

For small values of 𝜏𝐵 /𝜏𝑤 , as is seen from the results of the loop test experiment,

the mean velocity is approximated by the following linear expression.

𝐷 4
𝑢𝑚𝑒𝑎𝑛 = (𝜏𝑤 − 𝜏𝐵 ) (23)
8𝜇𝑝 3

By replacing 𝜏𝑤 in Eq. (23) with that from Eq. (6), the above equation can be re-

written and friction factor can be expressed in terms of values that may be experimentally

obtained i.e. Bingham yield stress, Bingham plastic viscosity and mean velocity.

32𝜇𝑝 𝑢𝑚𝑒𝑎𝑛 16𝜏𝐵 2𝐷


𝑓=( + ) (24)
𝐷2 2
3𝐷 𝜌𝑢𝑚𝑒𝑎𝑛

12
5.6. Darby-Melson
R. Darby and J. Melson developed a friction factor correlation by introducing a

modified Reynolds number to the Buckingham-Reiner equation. The model was initially

developed for Bingham plastic fluids with 𝜏𝐵 < 2Pa and 𝜇𝑝 < 0.2 Pas by modifying the

Buckingham-Reiner equation in terms of fanning friction factor by introducing a

modified Reynolds number.

16
𝑓= (25)
𝑅𝑒𝑚𝑜𝑑

6𝑅𝑒𝐵2
𝑅𝑒𝑚𝑜𝑑 = (26)
6𝑅𝑒𝐵 + 𝐻𝑒

5.7. Ihle and Tamburrino


C.F. Ihle and A. Tamburrino attempted to solve for the final expression of the

largest root of the Buckingham-Reiner equation. They introduced two parameters 𝑃 and

𝑄 to simplify Eq. (9) to Eq. (29) by taking Bingham number 𝐵=0.

𝐵 𝜏𝐵 𝐷
𝑃= ;𝐵 = (27)
𝑅𝑒𝐵 𝑢𝜇𝑝

2 𝐵
𝑄= ( + 2) (28)
𝑅𝑒𝐵 3

1 𝑃4 1 𝑃8 1 𝑃12
𝑓 = 4𝑄 − − − −⋯ (29)
12 𝑄 3 192 𝑄 7 9216 𝑄11

The first term on the right-hand side of Eq. (29) is analogous to that obtained by

dropping the fourth order term in Eq. (146. For small values of B, Eq. (30) is a good

approximation and for large values of B, Eq. (31) is a feasible solution.

1 8 9 𝐵4
𝑓= [16 + 𝐵 − ] ; 𝐵 = 𝑆𝑚𝑎𝑙𝑙 (30)
𝑅𝑒𝐵 3 32 (𝐵 + 6)3

2𝐵
𝑓= ; 𝐵 = 𝐿𝑎𝑟𝑔𝑒 (31)
𝑅𝑒𝐵

13
5.8. Hanks
The friction factor model developed by R.W. Hanks for the laminar flow of

Herschel-Bulkley fluids is shown below in Eq. (32). Eqs. (33) and (34) define the factor

𝜓 and dimensionless plug radius 𝜉0 . For Newtonian fluids, the friction factor is obtained

from Eq. (32) by taking 𝜓=1. For Bingham plastic fluids, 𝜓 and 𝜉0 are determined by

taking 𝑛=1. It is observed that when wall shear stress 𝜏𝑤 approaches the Bingham yield

stress 𝜏𝐵 , 𝜉0 →1 and 𝜓→0 resulting in infinitely high values of friction factor. In the

laminar range 𝜉0 must always be less than 1.

16
𝑓= (32)
𝜓𝑅𝑒𝐵
𝑛
𝑛 (1
(1 − 𝜉0 )2 2𝜉0 (1 − 𝜉0 ) 𝜉02
𝜓 = (1 + 3𝑛) − 𝜉0 )1+𝑛 [ + + ] (33)
(1 + 3𝑛) (1 + 2𝑛) (1 + 𝑛)

𝜏𝐵
𝜉0 = (34)
𝜏𝑤

6. Numerical Model
The numerical study of the backfill flow in the flow loop apparatus is performed

using the commercial computational fluid dynamics (CFD) solver ANSYS  Fluent. The

backfill fluid is considered a homogeneous single-phase fluid exhibiting non-Newtonian

Bingham plastic behaviour.

6.1. Domain and Mesh

A geometrical model is developed in ANSYS Workbench based on the

dimensions from Fig. 2. Only the portion of the loop test that has pressure gauges is

considered for numerical solving. The geometry is meshed using ANSYS Meshing with

a boundary layer mesh set to the O-grid on the pipe circumference towards the centre

14
ensuring finer mesh elements in the boundary. Six meshes with different degrees of

refinement were produced and the case with highest Reynolds number, (with lowest mass

concentration) was used to run the grid independence study. The results of the study are

shown in Fig. 3. The mesh with 3 million elements was selected for the numerical model

and is shown below in Figs. 4a and 4b.

6.2. Fluid Model


The mine backfill is assumed to be a single-phase fluid and the non-Newtonian

Bingham properties from loop test are used to model the fluid behaviour. A Bingham

fluid is one that requires a finite shearing stress to initiate motion and for which there

exists a linear relationship between the shearing stress greater than the initiating and the

resulting velocity gradient [38]. This minimum stress required is called critical yield

stress and it causes the fluid to flow as a viscous material with a finite viscosity. The

characteristic Bingham equation is displayed below.

𝜏 = 𝜏𝐵 + 𝜇 𝑃 𝛾 (35)

The apparent viscosity 𝑎𝑝𝑝 is obtained by dividing Eq. (35) by shear rate 𝛾

throughout.
𝜏 𝜏𝐵
𝑎𝑝𝑝 = = + 𝜇𝑃 (36)
𝛾 𝛾

For Bingham fluids, the apparent viscosity 𝑎𝑝𝑝 is infinite up until the applied

shear stress is below 𝜏𝐵 and beyond that it has a constant value 𝜇𝑃 . For numerical

modelling a small value of shear rate 0.001 is chosen for stitching up the discontinuity at

zero shear rate. During numerical solving, as the shear rate increases from 0 to 0.001,

𝑎𝑝𝑝 has a high value behaving almost as a solid. Beyond that, the Eq. (33) is obeyed.

15
6.3. Governing Equations
The numerical model solves the equations for mass and momentum conservation

that are defined below in Eqs. (37) and (38).

∇ ∙ 𝜌𝑢
⃗ =0 (37)

∇ ∙ (𝜌𝑢 ⃗ ) = −∇ ∙ 𝑝I + ∇ ∙ [𝑎𝑝𝑝 (∇𝑢


⃗ ×𝑢 ⃗ )𝑇 )] + 𝜌𝑔
⃗ + (∇𝑢 (38)

6.4. Boundary Conditions


The velocity profile at the inlet of the model is not known due to the pump

preceding the inlet in the actual set up. To counter this, a 2.0 m developing length is

included before the first pressure gauge. The average flow velocity is known, and a

uniform velocity profile is specified at the inlet surface. The walls have been specified

with no-slip condition, and outlet with gauge pressure equal to zero. The operating density

is set equal to the backfill density to ensure no hydrostatic pressure variation. This

facilitates the removal of gravitational head in the pressure variation in the pipe for

comparison with the experimental values.

6.5. Numerical methodology


The numerical setup for the Bingham model is verified against an analytical

solution of a two-dimensional axisymmetric pipe flow using the selected mesh density

for the two mines. The least squares cell-based gradient evaluation is used for its higher

accuracy and cheaper computation requirements over its node-based counterparts. Due to

the absence of strong body forces and stability issues, the second-order pressure scheme

is used. Different schemes for the pressure-velocity coupling and momentum were tested

to determine the most accurate velocity profile matching the analytical Bingham case.

The two pressure-velocity coupling algorithms tested were semi-implicit pressure linked

equation (SIMPLE) and pressure-implicit with splitting of operators (PISO). For spatial

16
discretization 1st order, 2nd order and 3rd order MUSCL (monotone upstream-centred

schemes for conservation laws) were tested. The velocity gradient from all the above

schemes were compared against the analytical results obtained from the below Eq. (40).

𝛥𝑃 2 𝑟 2
𝑢 = 𝜏𝐵 (𝑟 − 𝑅) + ( 𝑅 ) [1 − ( ) ] (40)
4𝜇𝑃 𝐿 𝑅

The results from the analyses are displayed below in Figs. 5a and 5b. The SIMPLE

2nd order produced the best agreement with the analytical solution for both mine samples.

Therefore, for all the remaining simulations were carried out using SIMPLE 2nd order

discretization scheme.

7. Test Results
7.1. Rheological Parameters from the Loop Test

The rheological parameters developed from the loop test experiment are presented

in Table 3. The experiment showed an increase in Bingham yield stress and Bingham

plastic viscosity with increasing solid concentrations. This phenomenon can be attributed

to the low mean particle and inter-particle distances in high solid concentration flows

resulting in strong interference effects that in-turn increase resistance to shearing [39].

7.2. Reynolds Number

The Bingham plastic viscosity determined through the loop test experiment is

input in Eq. (9) to get the Bingham Reynolds number and in Eq. (10) to get the

Generalized Bingham Reynolds number. A large deviation was seen between the two sets

of values. A parity plot between the two values displayed in Fig. 6 shows the

𝑅𝑒𝐺𝐸𝑁 𝐵𝐼𝑁𝐺𝐻𝐴𝑀 under predicts the Reynolds number in comparison to 𝑅𝑒𝐵 which results

in an over prediction of the friction factor. The friction factor at low velocities exceeded

1500 which is unrealistic. This confirms that the Generalized Bingham Reynolds number

17
proposed by K. Madlener, B. Frey and H.K. Ciezki is not suitable for high concentration

paste backfill. The Bingham Reynolds number in Eq. (9) is considered apt for this

application and is used henceforth in this study.

7.3. Friction Factor

Eight friction factor models from Eqs. (17) to (34) were compared against results

from CFD numerical simulation and loop test experiment conducted at three temperatures

16°C, 26°C and 35°C at low Reynolds numbers (<2300). The results are plotted in below

in Figs. 7a – 7f.

The Swamee & Aggarwal model as displayed in Eq. (18) showed the best

agreement with loop test data. The authors in their research noted less than 1% error in

their implicit formulation. The Jennifer Sinclair Curtis model as shown in Eq. (24) also

closely agrees with loop test data. Both models show better agreement with experimental

data at higher values of 𝑅𝑒𝐵 and start deviating marginally as 𝑅𝑒𝐵 reduces.

The Hagen-Poissuille model shown in Eq. (17) was initially derived for

Newtonian fluids and does not consider the Bingham yield stress. Despite not agreeing

with the loop test results, it is observed that this model agrees more closely than Hanks,

Ihle & Tamburrino, Danish & Kumar and Darby & Melson models in most cases putting

into question the validity of these models for high concentration paste backfill flow.

The Darby and Melson model as shown in Eq. (25) is very simple and was

originally developed for Bingham plastic fluids with very low yield stresses and plastic

viscosities (𝜏𝐵 < 2 Pa and 𝜇𝑝 < 0.2 Pas). The results from this model were very close to the

Danish & Kumar model as shown in Eq. (19). From their research, the authors concluded

their model had lower errors for turbulent flow (within 0.005%) in comparison to laminar

flows (within 5.2%). Both these models largely under predicts the friction factor.

18
The Ihle and Tamburrino model is divided into two cases; for small values of

Bingham number as shown in Eq. (30) and for large values of Bingham number as shown

in Eq. (31). In their study, the authors do not clearly prescribe what is classified as a small

or large Bingham number. From the loop test experiment, the Bingham number that were

recorded range from 6 to 117. Both models under predict the friction factor seen in the

loop test.

The Hanks model as shown in Eqs. (32) to (34) under predicts the friction factor.

The reason for this may be due to empirically obtaining the value of the factor 𝜓 and

dimensionless plug radius 𝜉0 which is a ratio of Bingham yield stress to wall shear stress.

At low Reynolds when 𝜏𝑤 approaches 𝜏𝐵 , 𝜉0 value approaches 1. This results in

unreasonably high friction factor. The over-reliance on 𝜉0 in this model makes it

unsuitable for low Reynold number flows.

The results from the CFD numerical simulation show very good agreement with

loop test data. The accuracy of the numerical simulation may further be improved by

accounting for the particle-particle interaction which is not incorporated in the current

model. For thick backfill slurries with high viscosity this phenomenon has a greater effect

at high Reynolds number. Another improvement can be done by accounting for the

temperature changes during the flow and how that may in turn affect the material

properties. The mine backfill which is in practice a multiphase flow (tailings and water)

is modelled as a single-phase fluid but with non-Newtonian Bingham properties. A more

advanced multiphase model will be considered in future to model paste backfill.

In summary, the results from this study as seen in Figs. 7a – 7f can be broadly

classified into two groups. The Swamee & Aggarwal, Jennifer Sinclair Curtis and the

CFD models show good agreement with each other and are far off from Danish & Kumar,

Hagen-Poisseuille, Darby & Melson, Ihle & Tamburrino and Hanks models.

19
7.4. Statistical Analysis
For better interpretation of the vast amount of data generated from each friction

factor model, statistical analysis techniques were employed. R2 or coefficient of

determination is a statistical measure of how close the data is to the fitted regression line.

It is the ratio of explained variation to total variation and lies between 0% and 100%.

Root mean square error (RMSE) is the standard deviation of residuals. It is a measure of

how spread out the predicted errors are. Mean absolute percent error (MAPE) is a measure

of the size of error in percentage terms. Percentage errors are scale-independent and range

from undefined to infinite. Symmetric mean absolute percent error (SMAPE) is a

variation of MAPE that ranges from -200 to 200. It overcomes the bias MAPE has

favouring predicted values that are lower than actual values. Two model selection criteria

were used in this study; Akaike Information Criterion (AIC) and Bayesian Information

Criterion (BIC). AIC and BIC are based on maximum likelihood estimates of the model

parameters. Table 4 below showcases the governing expression used in each statistical

model and its interpretation criterion. Tables 5 and 6 showcase the results obtained from

each statistical analysis for Mines A and B.

The coefficient of determination R2 gives high values above 0.97 for CFD,

Swamee & Aggarwal, Danish & Kumar, Jennifer Sinclair Curtis, Darby & Melson and

Ihle & Tamburino B=Large. The RMSE analysis shows lowest regression for Swamee &

Aggarwal (0.647) followed by CFD (0.685) and Jennifer Sinclair Curtis (0.966) while all

other models have values above 2.5. A similar trend is seen from MAPE and SMAPE

analyses where Swamee & Aggarwal has the lowest percentage error (3.984 and 3.860)

followed by Jennifer Sinclair Curtis and CFD. The model selection criteria AIC and BIC

gives the lowest values for Swamee & Aggarwal (116.810 and 116.486) followed by CFD

and Jennifer Sinclair Curtis while all other models have values above 200.

20
From the statistical analyses performed it can be concluded that Swamee &

Aggarwal model has the best fit with the experimental results obtained from the loop test.

The CFD model has the second-best fit based on RMSE, AIC and BIC while Jennifer

Sinclair Curtis has a better fit based on MAPE and SMAPE. It is important to note that

the R2 model shows inconsistent results by having high values for Danish & Kumar,

Darby & Melson and Ihle & Tamburrino B=Large which neither correlates with the

results from the other statistical models nor with the results in Figs. 7a – 7f. This indicates

that R2 is not a conclusive method to indicate a good fit.

7.5. In-situ mine data validation


The frictional pressure losses along the pipeline in Mine-A with total length of

more than three kilometres was measured using a PSI pill [40] in the mine. The pill was

introduced into the paste backfill pipeline at the inlet and collected at the discharge. The

data from the pill was extracted and compared against the frictional pressure loss

determined using the Swamee & Aggarwal friction factor model. The gravitational

pressure was removed from the PSI pill and Swamee & Aggarwal analyses using the

relation described in Eq. (41).

𝛥𝑃𝐹 = 𝛥𝑃 + (𝜌𝑔𝐿𝑉 ) (41)

A numerical model was prepared based on the pipeline schematic displayed in

Fig. 8a and the results were included in the comparison. Fig. 8b below shows the results

of this comparison and a good agreement is observed between the in-situ pill test results

numerical model and Swamee & Aggarwal friction factor model within an error

percentage of 5%. The numerical model showed better agreement towards the end of the

pipeline than the start. This comparison confirms the accuracy of the Swamee &

Aggarwal model in predicting the friction factor and the pressure losses in long distance

underground paste backfill pipelines. In the numerical model, an isothermal case is

21
adopted with the gravitational pressure decoupled, and the pressure loss is purely

frictional, which show the accuracy and validity of both numerical and friction factor

models for paste backfill application.

8. Conclusion

An experimental setup, called the loop test, was used to determine the pressure

gradients at different flow velocities for the flow of high concentration (67-75% by mass)

paste backfill samples prepared from tailings from two mines in Canada. The flow lies in

the laminar regime with Reynolds number < 2300. The friction factors for the flow were

obtained from the loop test experiment, numerical simulations and eight friction factor

correlations.

The Generalized Bingham Reynolds number was shown to under predict the

Reynolds number in comparison to loop test data and over predict the friction factor. It is

deemed incompatible for high concentration paste backfill flows. The Bingham Reynolds

number first proposed by Metzner and Reed is recommended.

Swamee & Aggarwal friction factor correlation was seen to best predict the

friction factor at the Reynolds number and temperature ranges tested in the loop test

experiment. The second closest agreement was recorded with the Jennifer Sinclair Curtis

model. The models proposed by Hagen-Poiseuille, Danish & Kumar, Darby & Melson,

Ihle & Tamburrino and Hanks predominantly under predict the friction factor. The results

from six statistical analyses confirm that Swamee & Aggarwal correlation best predicts

the friction factor from the experiment closely followed by Jennifer Sinclair Curtis model.

The CFD modelling of the flow loop system was done by considering the paste backfill

as a single-phase fluid with non-Newtonian Bingham properties. The numerical CFD

results show very good agreement with the loop test data. The CFD model can be further

22
improved by accounting for the particle-particle interaction and temperature dependence

of the paste backfill.

A comparison of frictional pressure losses measured in-situ in Mine-A using a pill

compared against the Swamee & Aggarwal correlation showed good agreement between

the two. Future work should be targeted at testing a wide range of mine tailings with

diverse mineral compositions and rheology. A friction factor correlation that

encompasses all paste backfill will be useful for engineers, designers and researchers in

underground mining applications.

This study shows that care should be taken in selecting a friction factor correlation

for non-Newtonian fluids. There exist numerous relationships derived for specific flow

regimes and fluid properties that remain valid only for the confines of the parameters

assumed while developing them.

Future work will focus on the two-fluid CFD model to capture the effects of

individual phases of water and tailings and their interactions. The current work will serve

as a pre-cursor to the two-fluid model which will account for particle treatment and

interphase momentum transfer.

Acknowledgement

The authors acknowledge the support of the Natural Sciences and Engineering

Research Council of Canada – Collaborative Research & Development (NSERC - CRD)

program. We also gratefully acknowledge Paterson & Cooke for the use of their loop test

facility to perform the test work. The third author acknowledges Mitacs Globalink

Program for supporting summer research internship at McGill University.

23
Nomenclature

𝜌 Density kg/m3

𝑃 Pressure Pa

𝑔 Acceleration due to gravity m/s2

𝐿 Unit length m

𝐿𝑉 Vertical length m

𝑆𝑔 Specific gravity -

𝛥𝑃 Pressure loss Pa

𝛥𝑃𝐹 Frictional pressure loss Pa

𝜏 Shear stress Pa

𝛾 Shear rate 1/s

𝛾′ Pseudo shear rate 1/s

𝑛 Flow behaviour index -

𝜏𝐵 Bingham yield stress Pa

𝜇𝑃 Bingham plastic viscosity Pas

𝑎𝑝𝑝 Apparent viscosity Pas

𝜇 Dynamic viscosity Pas

24
𝜏𝑤 Wall shear stress Pa

𝑢
⃗ Velocity vector m/s

𝑢 Flow velocity m/s

𝑢𝑚𝑒𝑎𝑛 Mean flow velocity m/s

𝐷 Pipe diameter m

𝑟, 𝑅 Pipe radius m

𝑅𝑒𝑛𝑒𝑤𝑡 Newtonian Reynolds number -

𝑅𝑒𝑔𝑒𝑛 Generalized Reynolds number -

𝑅𝑒𝐵 Bingham Reynolds number -

𝑅𝑒𝐺𝐸𝑁 𝐵𝐼𝑁𝐺𝐻𝐴𝑀 Generalized Bingham Reynolds number -

𝑅𝑒𝑚𝑜𝑑 Modified Reynolds number -

𝐻𝑒 Hedstrom number -

𝑚 Local exponential factor -

𝑓 Friction factor -

𝐾1 , 𝐾2 Constants -

𝐵 Bingham number -

𝑃, 𝑄 Parameters for solving -

25
𝜉0 Dimensionless plug radius -

𝜓 Factor for solving -

𝐼 Identity matrix -

𝐴𝑡 Actual value from experiment -

𝐹𝑡 Predicted value from friction factor model -

𝐹̅𝑡 Average response -

𝐹̂𝑡 Estimated regression line -

𝑁 Sample size -

𝑝 Number of parameters -

𝑅𝑆𝑆 Residual sum of squares -

References

[1] J. Barrett, M. Coulthard, P. Dight, Determination of fill stability, in: 12th Can.
Rock Mech. Symp., Canadian Institute of Mining, Metallurgy and Petroleum,
Sudbury, Ontario, 1978: pp. 85–91.

[2] B.G. Lottermoser, Recycling, Reuse and Rehabilitation of Mine Wastes, Elements.
7 (2011) 405 LP-410.
http://elements.geoscienceworld.org/content/7/6/405.abstract.

[3] R. Rankine, M. Pacheco, N. Sivakugan, Underground Mining with Backfills, Solis


and Rocks. 30 (2007) 93–101.

[4] D. Wu, Y. BG, L. YC, Pressure drop in loop pipe flow of fresh cemented coal
gangue-fly ash slurry: Experiment and simulation, Adv. Powder Technol. 26
(2015) 920–927. doi:10.1016/j.apt.2015.03.009.

[5] N. Cruz, Y. Peng, Rheology measurements for flotation slurries with high clay
contents – A critical review, Miner. Eng. 98 (2016) 137–150.
doi:https://doi.org/10.1016/j.mineng.2016.08.011.

26
[6] N. Gharib, B. Bharathan, L. Amiri, M. McGuinness, F.P. Hassani, A.P. Sasmito,
Flow characteristics and wear prediction of Herschel-Bulkley non-Newtonian
paste backfill in pipe elbows, Can. J. Chem. Eng. 95 (2017) 1181–1191.
doi:10.1002/cjce.22749.

[7] A. Rawat, S.N. Singh, V. Seshadri, Erosion wear studies on high concentration fly
ash slurries, Wear. 378–379 (2017) 114–125.
doi:https://doi.org/10.1016/j.wear.2017.02.039.

[8] D. Wu, M. Fall, S.J. Cai, Coupling temperature, cement hydration and rheological
behaviour of fresh cemented paste backfill, Miner. Eng. 42 (2013) 76–87.
doi:https://doi.org/10.1016/j.mineng.2012.11.011.

[9] K.J. Creber, M.F. Kermani, M. McGuinness, F.P. Hassani, In situ investigation of
mine backfill distribution system wear rates in Canadian mines, Int. J. Mining,
Reclam. Environ. 31 (2017) 426–438. doi:10.1080/17480930.2017.1339169.

[10] P.K. Senapati, B.K. Mishra, Design considerations for hydraulic backfilling with
coal combustion products (CCPs) at high solids concentrations, Powder Technol.
229 (2012) 119–125. doi:https://doi.org/10.1016/j.powtec.2012.06.018.

[11] J. Liu, R. Wang, F. Gao, J. Zhou, K. Cen, Rheology and thixotropic properties of
slurry fuel prepared using municipal wastewater sludge and coal, Chem. Eng. Sci.
76 (2012) 1–8. doi:https://doi.org/10.1016/j.ces.2012.04.010.

[12] K.M. Assefa, D. Kaushal, A comparative study of friction factor correlations for
high concentrate slurry flow in smooth pipes, J. Hydrol. Hydromechanics. 63
(2015) 13. doi:10.1515/johh-2015-0008.

[13] S. Chandel, V. Seshadri, S.N. Singh, Effect of Additive on Pressure Drop and
Rheological Characteristics of Fly Ash Slurry at High Concentration, Part. Sci.
Technol. 27 (2009) 271–284. doi:10.1080/02726350902922036.

[14] S. Chandel, S.N. Singh, V. Seshadri, Transportation of high concentration coal ash
slurries through pipelines, Part. Sci. Technol. 1 (2010) 1–9.

[15] K.C. Wilson, A.D. Thomas, A new analysis of the turbulent flow of non-newtonian
fluids, Can. J. Chem. Eng. 63 (1985) 539–546. doi:10.1002/cjce.5450630403.

[16] F.A. Morrison, Data correlation for friction factor in smooth pipes, Houghton,
Michigan, 2013.

[17] D. Wu, B. Yang, Y. Liu, Transportability and pressure drop of fresh cemented coal
gangue-fly ash backfill (CGFB) slurry in pipe loop, Powder Technol. 284 (2015)
218–224. doi:https://doi.org/10.1016/j.powtec.2015.06.072.

[18] P.K. Swamee, N. Aggarwal, Explicit equations for laminar flow of Bingham
plastic fluids, J. Pet. Sci. Eng. 76 (2011) 178–184.
doi:https://doi.org/10.1016/j.petrol.2011.01.015.

[19] C. Qi, Q. Chen, A. Fourie, J. Zhao, Q. Zhang, Pressure drop in pipe flow of
cemented paste backfill: Experimental and modeling study, Powder Technol.

27
(2018). doi:https://doi.org/10.1016/j.powtec.2018.03.070.

[20] K. Madlener, B. Frey, H.K. Ciezki, Generalized reynolds number for non-
newtonian fluids, EUCASS Proc. Ser. – Adv. Aerosp. Sci. 1 (2009) 237–250.
https://doi.org/10.1051/eucass/200901237.

[21] R.P. Chhabra, Non-Newtonian Fluids: An Introduction BT - Rheology of


Complex Fluids, in: J.M. Krishnan, A.P. Deshpande, P.B.S. Kumar (Eds.),
Springer New York, New York, NY, 2010: pp. 3–34. doi:10.1007/978-1-4419-
6494-6_1.

[22] C.W. Bakker, C.J. Meyer, D.A. Deglon, Numerical modelling of non-Newtonian
slurry in a mechanical flotation cell, Miner. Eng. 22 (2009) 944–950.
doi:https://doi.org/10.1016/j.mineng.2009.03.016.

[23] V.R. Gopala, J.-A. Lycklama à Nijeholt, P. Bakker, B. Haverkate, Development


and validation of a CFD model predicting the backfill process of a nuclear waste
gallery, Nucl. Eng. Des. 241 (2011) 2508–2518.
doi:https://doi.org/10.1016/j.nucengdes.2011.04.021.

[24] F. Jiang, P. Zhao, G. Qi, N. Li, Y. Bian, H. Li, T. Jiang, X. Li, C. Yu, Pressure
drop in horizontal multi-tube liquid–solid circulating fluidized bed, Powder
Technol. (2018). doi:https://doi.org/10.1016/j.powtec.2018.04.003.

[25] A.B. Metzner, J.C. Reed, Flow of non-newtonian fluids—correlation of the


laminar, transition, and turbulent-flow regions, AIChE J. 1 (1955) 434–440.
doi:10.1002/aic.690010409.

[26] E.J. Garcia, J.F. Steffe, Comparison of friction factor equations for non-Newtonian
fluids in pipe flow, J. Food Process Eng. 9 (1986) 93–120. doi:10.1111/j.1745-
4530.1986.tb00120.x.

[27] S.P. Sutera, R. Skalak, The History of Poiseuille’s Law, Annu. Rev. Fluid Mech.
25 (1993) 1–20. doi:10.1146/annurev.fl.25.010193.000245.

[28] H. Blasius, Das Aehnlichkeitsgesetz bei Reibungsvorgangen in Flussigkeiten. [The


Law of Similarity in Friction Processes in Liquids], VDI Mitt. Forschungsarb. 131
(1913) 1–39.

[29] E. Buckingham, On plastic flow through capillary tubes, in: ASTM, 1921: pp.
1154–1161.

[30] C.F. Colebrook, Turbulent flow in pipes, with particular reference to the transition
region between the smooth and pipe laws, J. Inst. Civ. Eng. 11 (1939) 133–156.
doi:10.1680/ijoti.1939.13150.

[31] L. Prandtl, The mechanics of viscous fluids, Durand, W.F. Aerodyn. Theory. 3
(1935) 34–208.

[32] R. Darby, J. Melson, How to predict the friction factor for flow of Bingham
plastics, Chem. Eng. 88 (1981) 59–61.

28
[33] P.T. Slatter, The turbulent flow of non-Newtonian slurries in pipe, in: 8th Conf.
Transp. Sediment. Solid Part., Prague, 1995.

[34] M. Danish, S. Kumar, S. Kumar, Approximate explicit analytical expressions of


friction factor for flow of Bingham fluids in smooth pipes using Adomian
decomposition method, Commun. Nonlinear Sci. Numer. Simul. 16 (2011) 239–
251. doi:https://doi.org/10.1016/j.cnsns.2010.03.013.

[35] J.S. Curtis, Slurry Transport, in: M.J. Rhodes (Ed.), Introd. to Part. Technol., 2nd
ed., 2008: pp. 91–116. doi:10.1002/9780470727102.

[36] C.F. Ihle, A. Tamburrino, A note on the Buckingham equation, Can. J. Chem. Eng.
90 (2012) 944–945. doi:10.1002/cjce.20606.

[37] R.W. Hanks, Low Reynolds number turbulent pipeline flow of


pseudohomogeneous slurries, in: Fifth Int. Conf. Hydraul. Transp. Solids Pipes
(Hydrotransport 5), BHRA Fluid Engineering, Cranfield, Bedford, England,
Hannover, Germany, 1978: pp. C2-23-34.

[38] G.W. Govier, K. Aziz, The flow of complex mixtures in pipes, Van Nostrand
Reinhold, 1972.

[39] S. Chandel, S.N. Singh, V. Seshadri, Deposition characteristics of coal ash slurries
at higher concentrations, Adv. Powder Technol. 20 (2009) 383–389.
doi:https://doi.org/10.1016/j.apt.2009.06.004.

[40] Camiro. (2004). "Know Your Pipeline - PSI Pill (Parr Innovations) " Retrieved
August 30, 2011, from http://www.camiro.org/Mining_Products.htm

29
Table 1
Mine-A and B tailings composition.

Oxide Mine-A Mine-B Oxide Mine-A Mine-B

CO2 - 0.88% K2O 3.46% 2.70%

Na2O 3.32% 1.50% CaO 5.67% 6.52%

MgO 3.26% 3.97% TiO2 0.57% 0.64%

Al2O3 18.64% 17.67% MnO 0.15% 0.13%

SiO2 54.79% 49.02% Fe2O3 6.75% 9.15%

SO3 2.83% 7.29% ZnO - 0.15%

30
Table 2
Mine-A and B tailings characterization.

Parameter Mine-A Mine-B

d80 60 µm 20 µm

d50 25 µm 10 µm

d20 8 µm 7 µm

% < 20 µm 45 % 80 %

Specific Gravity 2.74 3.40

31
Table 3
Rheological parameters developed from loop test data for Mine-A and Mine-B.
Mine-A Mine-B

Bingham Bingham Bingham Bingham


Solid Solid
Temp. Yield Plastic Density Temp. Yield Plastic Density
Fraction 3
Fraction
(°C) Stress Viscosity (kg/m ) (°C) Stress Viscosity (kg/m3)
(%) (%)
(Pa) (Pa s) (Pa) (Pa s)

74.14 176.18 0.5457 1886 72.25 315.04 2.6003 1895

72.44 76.53 0.2860 1848 70.05 146.98 0.8863 1845


16 16
71.84 52.79 0.2917 1835 69.25 105.82 0.3905 1827

71.74 41.00 0.2085 1833 66.86 71.16 0.1514 1776

75.33 306.79 1.0255 1911 71.54 239.05 1.0349 1877

75.03 213.81 0.9863 1904 70.94 162.75 0.8503 1863

26 74.53 240.28 0.2637 1893 26 69.25 126.69 0.2053 1825

74.33 171.10 0.3104 1888 68.56 91.43 0.1736 1810

74.03 149.93 0.1795 1881 67.56 68.68 0.0993 1789

73.90 240.55 0.5205 1870 74.13 269.82 1.4994 1936

73.50 161.55 0.4890 1861 73.44 260.24 1.4363 1919

35 72.90 167.99 0.0869 1848 35 72.75 246.85 1.2916 1902

71.80 99.86 0.0780 1825 70.57 148.77 0.4983 1852

70.20 42.02 0.0761 1792 69.87 148.55 0.1499 1836

32
Table 4
Governing expression of each statistical model and its interpretations.

Model Governing expression Interpretation

Coefficient of ∑𝑁 ̅ ̂ 2
𝑖=1(𝐹𝑡 − 𝐹𝑡 )
The closer the value is to 1,
2
𝑅 = 𝑁
determination R2 ∑𝑖=1(𝐹𝑡 − 𝐹̅𝑡 )2 the better the model

Root Mean Square Error The smaller the errors, the


∑𝑁
𝑖=1(𝐹𝑡 − 𝐴𝑡 )
2
𝑅𝑀𝑆𝐸 = √
(RMSE) 𝑁 better the model

Mean Absolute Percent 𝑁 The smaller the percentage


100 |𝐹𝑡 − 𝐴𝑡 |
𝑀𝐴𝑃𝐸 = ∑
Error (MAPE) 𝑁 𝐴𝑡 error, the better the model
𝑖=1

Symmetric Mean Absolute 𝑁 The smaller the percentage


200 |𝐹𝑡 − 𝐴𝑡 |
𝑆𝑀𝐴𝑃𝐸 = ∑
Percent Error (SMAPE) 𝑁 |𝐹𝑡 + 𝐴𝑡 | error, the better the model
𝑖=1

𝑅𝑆𝑆
Akaike Information 𝐴𝐼𝐶 = 𝑁 + 𝑁𝑙𝑜𝑔(2𝜋) + 𝑁𝑙𝑜𝑔 ( ) The smaller the value, the
𝑁
Criterion (AIC) better the model
+ 2(𝑝 + 1)

𝑅𝑆𝑆
Bayesian Information 𝐵𝐼𝐶 = 𝑁 + 𝑁𝑙𝑜𝑔(2𝜋) + 𝑁𝑙𝑜𝑔 ( ) The smaller the value, the
𝑁
Criterion (BIC) better the model
+ (𝑙𝑜𝑔𝑁)(𝑝 + 1)

33
Table 5
Statistical analysis of friction factor models against loop test experiment for Mine-A.

Model R2 RMSE MAPE SMAPE AIC BIC

CFD 0.978 0.685 15.926 17.534 120.666 120.342

Hagen-Poisseuille 0.767 3.125 77.784 130.109 223.459 223.135

Swamee & Aggarwal 0.975 0.647 3.984 3.860 116.810 116.486

Danish & Kumar 0.975 2.687 74.724 119.347 213.215 212.891

Jennifer Sinclair Curtis 0.971 0.966 8.647 8.010 143.938 143.614

Darby & Melson 0.971 2.597 73.093 115.301 210.918 210.595

Ihle & Tamburrino


0.730 2.570 71.104 112.309 210.201 209.878
B=Small

Ihle & Tamburrino


0.952 2.982 83.986 145.084 220.271 219.947
B=Large

Hanks 0.431 2.806 72.811 116.606 216.166 215.843

34
Table 6
Statistical analysis of friction factor models against loop test experiment for Mine-B.

Model R2 RMSE MAPE SMAPE AIC BIC

CFD 0.987 0.958 12.810 14.011 132.755 132.327

Hagen-Poisseuille 0.787 4.521 71.323 113.901 229.829 229.401

Swamee & Aggarwal 0.988 0.617 2.807 2.770 105.271 104.843

Danish & Kumar 0.988 4.298 74.926 119.839 226.655 226.227

Jennifer Sinclair Curtis 0.983 0.922 5.829 5.529 130.382 129.954

Darby & Melson 0.983 4.208 73.796 117.007 225.332 224.904

Ihle & Tamburrino


0.987 4.280 74.712 119.298 226.392 225.964
B=Small

Ihle & Tamburrino


0.946 4.858 85.724 150.312 234.319 233.891
B=Large

Hanks 0.455 4.050 74.674 117.352 222.936 222.508

35
Figures

Fig. 1

Fig. 2

36
Fig. 3

Fig. 4a

37
Fig. 4b

Fig. 5a

38
Fig. 5b

Fig. 6

39
Fig. 7a

Fig. 7b

40
Fig. 7c

Fig. 7d

41
Fig. 7e

Fig. 7f

42
Fig. 8a

Fig. 8b

43
Figure Captions
Fig. 1. Mine-A and B particle size distribution.
Fig. 2. Schematic arrangement of the flow loop test experimental apparatus (All
dimensions are in mm).
Fig. 3. Grid independence study.
Fig. 4. Loop geometry used in CFD model; (a) computational domain and (b) mesh
profile.
Fig. 5. Numerical model verification against various discretization schemes for (a) Mine-
A; and (b) Mine-B.
Fig. 6. Bingham Reynolds number – Generalized Bingham Reynolds number parity plot.
Fig. 7. Comparison of friction factor models for Mine-A at (a) 16°C; (b) 26°C; (c) 35°C;
and for Mine-B (d) 16°C; (e) 26°C; and (f) 35°C.
Fig. 8. (a) Schematic diagram of the mine-field pipeline used in CFD model; and (b)
Comparison of pressure reading between in-situ mine data, Swamee & Aggarwal friction
factor model and CFD model.

44

View publication stats

Das könnte Ihnen auch gefallen