
Flotation is a unit operation extensively used in the recovery of valuable minerals in mineral processing and related applications. Essential insight to the hydrodynamics of a flotation column can be obtained by studying just two phases: gas and fluid. To this end, the approach based on the drift-flux theory, proposed in similar form by several authors, is reformulated as a one-dimensional non-linear conservation law with a multiply discontinuous flux. The unknown is the gas volume fraction as a function of height and time, and the flux function depends discontinuously on spatial position due to several feed inlets. The resulting model is similar, but not equivalent, to previously studied clarifier-thickener models for solid-liquid separation and therefore adds a new real-world application to the field of conservation laws with discontinuous flux. Steady-state solutions are studied in detail, including their construction by applying an appropriate entropy condition across each flux discontinuity. This analysis leads to operating charts and tables collecting all possible steady states along with some necessary conditions for their feasibility in each case. Numerical experiments show that the transient model recovers the steady states, depending on the feed rates of the different inlets.
Citation: Raimund Bürger, Stefan Diehl, María Carmen Martí. A conservation law with multiply discontinuous flux modelling a flotation column[J]. Networks and Heterogeneous Media, 2018, 13(2): 339-371. doi: 10.3934/nhm.2018015
[1] | Raimund Bürger, Stefan Diehl, María Carmen Martí . A conservation law with multiply discontinuous flux modelling a flotation column. Networks and Heterogeneous Media, 2018, 13(2): 339-371. doi: 10.3934/nhm.2018015 |
[2] | Raimund Bürger, Stefan Diehl, M. Carmen Martí, Yolanda Vásquez . A difference scheme for a triangular system of conservation laws with discontinuous flux modeling three-phase flows. Networks and Heterogeneous Media, 2023, 18(1): 140-190. doi: 10.3934/nhm.2023006 |
[3] | Mauro Garavello, Roberto Natalini, Benedetto Piccoli, Andrea Terracina . Conservation laws with discontinuous flux. Networks and Heterogeneous Media, 2007, 2(1): 159-179. doi: 10.3934/nhm.2007.2.159 |
[4] | Christophe Chalons, Paola Goatin, Nicolas Seguin . General constrained conservation laws. Application to pedestrian flow modeling. Networks and Heterogeneous Media, 2013, 8(2): 433-463. doi: 10.3934/nhm.2013.8.433 |
[5] | Raimund Bürger, Kenneth H. Karlsen, John D. Towers . On some difference schemes and entropy conditions for a class of multi-species kinematic flow models with discontinuous flux. Networks and Heterogeneous Media, 2010, 5(3): 461-485. doi: 10.3934/nhm.2010.5.461 |
[6] | Raimund Bürger, Christophe Chalons, Rafael Ordoñez, Luis Miguel Villada . A multiclass Lighthill-Whitham-Richards traffic model with a discontinuous velocity function. Networks and Heterogeneous Media, 2021, 16(2): 187-219. doi: 10.3934/nhm.2021004 |
[7] | Clément Cancès . On the effects of discontinuous capillarities for immiscible two-phase flows in porous media made of several rock-types. Networks and Heterogeneous Media, 2010, 5(3): 635-647. doi: 10.3934/nhm.2010.5.635 |
[8] | Felisia Angela Chiarello, Giuseppe Maria Coclite . Nonlocal scalar conservation laws with discontinuous flux. Networks and Heterogeneous Media, 2023, 18(1): 380-398. doi: 10.3934/nhm.2023015 |
[9] | Boris Andreianov, Kenneth H. Karlsen, Nils H. Risebro . On vanishing viscosity approximation of conservation laws with discontinuous flux. Networks and Heterogeneous Media, 2010, 5(3): 617-633. doi: 10.3934/nhm.2010.5.617 |
[10] | Wen Shen . Traveling wave profiles for a Follow-the-Leader model for traffic flow with rough road condition. Networks and Heterogeneous Media, 2018, 13(3): 449-478. doi: 10.3934/nhm.2018020 |
Flotation is a unit operation extensively used in the recovery of valuable minerals in mineral processing and related applications. Essential insight to the hydrodynamics of a flotation column can be obtained by studying just two phases: gas and fluid. To this end, the approach based on the drift-flux theory, proposed in similar form by several authors, is reformulated as a one-dimensional non-linear conservation law with a multiply discontinuous flux. The unknown is the gas volume fraction as a function of height and time, and the flux function depends discontinuously on spatial position due to several feed inlets. The resulting model is similar, but not equivalent, to previously studied clarifier-thickener models for solid-liquid separation and therefore adds a new real-world application to the field of conservation laws with discontinuous flux. Steady-state solutions are studied in detail, including their construction by applying an appropriate entropy condition across each flux discontinuity. This analysis leads to operating charts and tables collecting all possible steady states along with some necessary conditions for their feasibility in each case. Numerical experiments show that the transient model recovers the steady states, depending on the feed rates of the different inlets.
Flotation is a unit operation that is extensively used in the recovery of valuable minerals and coals in mineral processing but also in many other applications in environmental and chemical engineering [12,21,33,36,42]. It is a physico-chemical separation process that utilizes the difference in surface properties of the valuable hydrophobic minerals and the unwanted hydrophilic gangue material. The theory of froth flotation is complex and involves three phases (solids, water, and froth or gas) with many subprocesses [42]. The principle of the conventional flotation process is roughly as follows. Gas is introduced close to the bottom of a flotation column (see Figure 1), and the bubbles generated then rise upward throughout the pulp that contains hydrophobic and hydrophilic solid particles. The hydrophobic particles in the pulp attach to the bubbles. Since the overall density of the bubble-particle aggregates is less than that of the medium, the aggregates then float to the top of the column, where the desired product, the foam or froth carrying the valuable material (the concentrate in mining) is removed, usually through a launder. Additionally, close to the top wash water is injected to assist with the rejection of entrained impurities [39] and to increase the froth stability and improve recovery [31,21]. Once the hydrophobic particles have attached to the air bubbles, flotation can be considered as a separation between relatively large low-density entities, called air bubbles, and a suspension of liquid and gangue.Consequently, flotation can be described as a gas-liquid separation process by buoyancy analogous to the solid-liquid separation by gravity sedimentation in clarifier-thickeners [8,10,15].
Well-established spatially one-dimensional models of clarifier-thickeners can be formulated as a scalar conservation law for the local solids concentration as a function of depth and time, where the flux is discontinuous as a function of spatial position due to upward- and downward-directed bulk flows, transitions to overflow and underflow transport, and a singular source term marking the feed [8,10,15].Clarifier-thickener models have motivated in part the mathematical research on conservation laws with discontinuous flux [3,4,6,10,14,15,16,17,18,19,20].
It is the purpose of this paper to formulate, partially analyze, solve for steady states, and numerically simulate a related model for a flotation column, where we follow [13] and limit ourselves to a one-dimensional two-phase system of gas bubbles dispersed in a fluid, or rather a suspension of liquid and gangue. Hence, we do not model any sedimentation of solid particles in the suspension. The final form of the model (cf. Figure 1) is the conservation law with multiply discontinuous flux
$
∂ϕ∂t+∂∂zF(z,t,ϕ)=∑S∈{G,F,W}qS(t)ϕS(t)δ(z−zS),z∈R,t>0,
$
|
(1.1) |
where
$
ϕ(z,0)=ϕ0(z),z∈R.
$
|
(1.2) |
While the time-dependent partial differential equation (PDE) (1.1) describes transient variations of
The main outcomes of this work are to a classification of all steady-state solutions by means of diagrams and tables, and numerical simulations of dynamic behaviour. The variety of real-world applications of conservation laws with discontinuous flux is hereby widened to include flotation.
Our model formulation is based on the description of a flotation column by Stevenson et al. [38], Dickinson and Galvin [13], and Galvin and Dickinson [22] that is based on algebraic expressions for the gas and liquid fluxes, velocities and volume fractions. The description of one-dimensional two-phase flows based on the continuity equations for both phases and closed by defining a relative flux, or drift flux, between both phases as a function of volume fraction was introduced by Wallis [40], as is elaborated in [35]. Treatments that invoke this drift-flux analysis to describe flotation processes include [25,27,31,39,43,44]. However, all these works utilize these variables for steady-state analyses, but do not incorporate the drift-flux variables into one solvable PDE model for transient simulations, which is precisely the main contribution of the present paper.
As stated above, the theory of conservation laws with discontinuous flux has seen a vast amount of interest in recent years, where the typical model equation is
$∂ϕ∂t+∂∂zF(z,ϕ)=0,F(z,ϕ)={f(ϕ)forz>0,g(ϕ)forz<0 $
|
or equivalently, in terms of the Heaviside step function
$ \frac{\partial \phi}{\partial t} + \frac{\partial }{\partial z}\bigl( H(z)f(\phi)+(1-H(z))g(\phi)\bigr) = 0. $ | (1.3) |
The basic difficulty associated with (1.3) is as follows. Suppose, for simplicity, that
$
f(ϕ+(t))=g(ϕ−(t))for a.e.t>0.
$
|
(1.4) |
This single equation does not define the two traces uniquely and one needs to specify a selection principle or jump entropy condition to single out pairs that besides satisfying (1.4) are admissible. This selection principle usually depends on the particular physical reality (1.3) is supposed to model. For instance, applications of (1.3) also include traffic flow with discontinuously changing road surface conditions, ion etching, two-phase flow in heterogeneous porous media, and medical applications (see [5], [24,Ch. 8], and [28] for overviews and references). We use here the admissibility condition from [14], which has proved to be the natural one for the related problem of continuous sedimentation [15]. Furthermore, its generalization [18] to the case of a scalar convection-diffusion equation with spatial discontinuity in both the flux and diffusion functions implies the physically relevant solution in the case of the well-established model of continuous sedimentation with compression [10]. As is stated in [24,p. 426], there are different "recipes" to select unique solutions of the Riemann problem of (1.3), and all of them eventually lead to uniqueness of the initial value problem for (1.3), according to the unified treatment in [1].
In Section 2, we derive the model equations for the local fraction of gas bubbles, detailing the definition of the flux density functions in each zone of the spatial domain and the treatment of the feed inlets. Some notation necessary for the description of the steady-state solutions is also introduced. In Section 3, we focus on the characterization of all possible steady states for the flotation model previously defined, providing a detailed study of the derivation process at the different spatial discontinuities introduced by the feed inlets. Some steady states exist only under certain conditions on the injection rates and to get an overview of all possibilities, we present operating charts and tables for the categorization of all steady states. In Section 4, we briefly review the numerical method to approximately solve the flotation model. Some numerical examples are provided in Section 5 and, finally, we present some conclusions in Section 6.
Assume that we consider a region of space that is free of sources and sinks, that
$
∂ϕ∂t+∇⋅(ϕvg)=0,∂(1−ϕ)∂t+∇⋅((1−ϕ)vf)=0,
$
|
(2.1) |
where we assume that the gas bubbles are incompressible and do not coalesce. Then, defining the volume average velocity, or bulk flux of the suspension,
$ \boldsymbol{q} : = \phi \boldsymbol{v}_{\rm{g}} + (1-\phi) \boldsymbol{v}_{\rm{f}} $ | (2.2) |
and the gas-fluid relative velocity
$∂ϕ∂t+∇⋅(ϕq+ϕ(1−ϕ)vr)=0,∇⋅q=0. $
|
It is assumed that
$
jg(ϕ):=ϕ(1−ϕ)vr(ϕ)=ϕvtermV(ϕ).
$
|
(2.3) |
The drift-flux function
Furthermore, in (2.3),
$
V(ϕ)=(1−ϕ)nRZ,nRZ≥0.
$
|
(2.4) |
The maximum possible volume fraction of bubbles is
Finally, in one space dimension (in the
$
∂ϕ∂t+∂∂zj(ϕ,t)=0,where j(ϕ,t):=q(t)ϕ+jg(ϕ).
$
|
(2.5) |
Hence,
It is assumed that the unit has a constant cross-sectional area
The spatially piecewise constant bulk velocity
Starting from the bottom of the vessel, we have
$
q(z,t):={q1=−qUforz<zG,q2=qG−qUforzzG≤z<zF,q3=qG+qF−qUforzzF≤z<zW,q4=qG+qF+qW−qUforzz≥zW.
$
|
(2.6) |
Hence, we always have
We denote the intervals
We assume that in the effluent and underflow zones, the gas and the fluid move at the same velocity, so we set
$
F(z,t,ϕ):={jU(ϕ,t):=q1(t)ϕforz<zU,j1(ϕ,t):=q1(t)ϕ+jg(ϕ)forzU≤z<zG,j2(ϕ,t):=q2(t)ϕ+jg(ϕ)forzG≤z<zF,j3(ϕ,t):=q3(t)ϕ+jg(ϕ)forzF≤z<zW,j4(ϕ,t):=q4(t)ϕ+jg(ϕ)forzW≤z<zE,jE(ϕ,t):=q4(t)ϕforz≥zE.
$
|
(2.7) |
The conservation law (2.5) is completed by including the feed of material at levels
Within each zone, the governing equation (1.1), (2.7) reduces to (2.5), and we consider the Cauchy problem of this equation. A piecewise smooth function
$ z_{\rm d}'(t) = S(\phi^{+}, \phi^{-}): = \frac{f(\phi^{+})- f(\phi^{-})}{\phi^{+}-\phi^{-}}, $ | (2.8) |
and the jump entropy condition
$
S(u,ϕ−)≥S(ϕ+,ϕ−)for allubetweenϕ+andϕ−.
$
|
(2.9) |
It is well known that entropy solutions in the sense of Oleinik [30] are also the unique entropy solutions in the sense of Kružkov-type [26] integral inequalities (cf., e.g., [24]). On the other hand and as mentioned in Section 1.2, at the five spatial discontinuities of problem (1.1), a generalized entropy is needed [14,18]. Since we only construct steady-state solutions, we review that condition in Section 3.1 for this purpose, which means less notation than for the dynamic case.
We assume that the drift-flux function
Since we will mostly refer to steady-state situations, we often supress the dependence on
$ˉq:=−j′g(1),ˉˉq:=−j′g(ϕinfl), $
|
which are the bulk velocities such that the slope of
$ \bar q = -j_{\rm{g}}'(1) = 0, $ | (2.10) |
in accordance with the common Richardson-Zaki function (2.4). For intermediate values of
$
\phi_{k{\rm{M}}}: = {1ifqk≤ˉq,((jg|(ϕinfl,1))′)−1(−qk)ifˉq<qk<ˉˉq,ϕinflif qk≥ˉˉq.
$
|
Given
$ j_k(\phi_{k{\rm{m}}}) = j_k(\phi_{k{\rm{M}}}),\;\;\; 0\le\phi_{k{\rm{m}}}\le\phi_{\rm infl}. $ |
For
In some instances it is convenient to write out the dependence on
Lemma 2.1. The following properties hold:
$ddqkjk(ϕkM(qk);qk)=ϕkM(qk),ddqkjk(ϕMk(qk);qk)=ϕMk(qk). $
|
In order to completely describe all possible steady states of model (1.1) (under the assumptions of Section 2.5), we here extract from the theory of conservation laws with discontinuous flux function [14,18] what is necessary to construct steady-state solutions in a neighbourhood of a spatial discontinuity.
To start with, we investigate the case when the steady-state solution is constant
We consider the conservation law with discontinuous flux function (1.3). The equation should be interpreted in the weak sense and we seek steady-state solutions of the form
$ϕ(z)={ϕ−ifz<0,ϕ+ifz>0, $
|
where
$ˆf(ϕ;ϕ+):={minv∈[ϕ,ϕ+]f(v)ifϕ≤ϕ+,maxv∈[ϕ+,ϕ]f(v)ifϕ>ϕ+,ˇg(ϕ;ϕ−):={maxv∈[ϕ,ϕ−]g(v)ifϕ≤ϕ−,minv∈[ϕ−,ϕ]g(v)ifϕ>ϕ−}=ˆg(ϕ−;ϕ). $
|
Since
We define the set of possible
$ˉΦ=ˉΦ(ϕ+,ϕ−):={ϕ∈[0,1]:ˆf(ϕ;ϕ+)=ˇg(ϕ;ϕ−)} $
|
and the corresponding unique flux value
$ {\hat f}(\phi_+;\phi_+) = \eta(\phi_+, \phi_-) = {\check g}(\phi_-;\phi_-), $ | (3.1) |
where we note that
In a neighbourhood of
$∂ϕ∂t+∂∂zF(z,t,ϕ)=0, $
|
where
$F(z,t,ϕ)={jU(ϕ,t)=q1(t)ϕifz<zU,j1(ϕ,t)=q1(t)ϕ+jg(ϕ)ifz>zU. $
|
We now suppress the time dependence and seek possible constant solutions
$ {\hat{\jmath}}(\phi_1;\phi_1) = \eta(\phi_1, \phi_{\rm U}) = {\hat{\jmath}}_U(\phi_{\rm U};\phi_{\rm U}), $ | (3.2) |
where
(a)
(b)
We conclude this subsection by stating the possible steady-state values in the underflow and the first zone:
$ \phi_{\rm{U}}\in[0, 1], \;\;\;\phi_1\in\{0\}\cup[\phi_{1{\rm{Z}}}, 1]. $ | (3.3) |
Here,
$ \phi_4\in[0, \phi_{4{\rm{m}}}] \cup [\phi_{4{\rm{M}}}, 1],\;\;\; \phi_{\rm{E}}\in[0, 1]. $ | (3.4) |
In a neighbourhood of
$∂ϕ∂t+∂∂z((1−H(z−zG))j1(ϕ,t)+H(z−zG)j2(ϕ,t))=qG(t)ϕG(t)δ(z−zG), $
|
which, with
$∂ϕ∂t+∂∂z((1−H(z−zG))(j1(ϕ,t)+qG(t))+H(z−zG)j2(ϕ,t))=0. $
|
The flux functions to consider for the steady-state coupling are thus
$
j1(ϕ)+qG=jg(ϕ)−qUϕ+qGforz<zG,j2(ϕ)=jg(ϕ)+(qG−qU)ϕforz>zG,
$
|
which intersect only at
$ {\hat \jmath}_2(\phi_2;\phi_2) = \eta(\phi_1, \phi_2) = {\check \jmath}_1(\phi_1;\phi_1)+q_{\rm{G}}. $ |
We have
Remark 1. In the division into subcases further on, we will sometimes let such overlap in the following way. Instead of having disjoint intervals defining two subcases, e.g.,
1. Case G1.
(a)
$ j_2(\phi_2^{\rm{M}})\geq q_{\rm{G}}. $ | (G) |
(b)
(c)
(d)
(e)
(f)
2. Case G2.
(a)
(b)
(c)
(d)
We derive the possible constant steady states in zones 2 and 3 considering their coupling at
$ j_2(\phi) = q_2\phi + j_g(\phi), \;\;\;z < z_{\rm{F}}, \\ j_3(\phi) = q_3\phi + j_g(\phi), \;\;\;z > z_{\rm{F}}, $ |
where
$ {\hat \jmath}_3(\phi_3;\phi_3) = \eta(\phi_3, \phi_2) = {\check{\jmath}}_2(\phi_2;\phi_2). $ | (3.5) |
We now study the possible intersections between
1. Case F1.
(a)
(b)
(c)
$ j_2(1)\geq j_2(\phi_2^{\rm{M}}). $ | (FⅠ) |
(d)
$ j_2(\phi_2^{\rm{M}}) \geq j_3(\phi_{3{\rm{M}}}). $ | (FⅡ) |
(e)
(f)
$ j_2(1)\geq j_3(\phi_{3{\rm{M}}}). $ | (FⅢ) |
(g)
(h)
(i)
2. Case F2.
(a)
(b)
(c)
(d)
(e)
(f)
3. Case F3.
(a)
(b)
(c)
(d)
Since
1. Case W1.
(a)
(b)
(c)
(d)
$ j_3(\phi_3^{\rm{M}}) \geq j_4(\phi_{4{\rm{M}}}). $ | (WⅠ) |
(e)
(f)
$ j_3(1) \geq j_4(\phi_{4{\rm{M}}}). $ | (WⅡ) |
2. Case W2.
The different necessary conditions on the fluxes that appear in the derivation of possible steady states can be visualized in operating charts. These are two-dimensional diagrams involving the bulk velocities at an injection point.
Condition (G) can be written as
$ q_{\rm{G}}\leq j_2(\phi_2^{\rm{M}}(q_2);q_2), $ | (G) |
where the dependence on
$
\left\{qU=j2(ϕM2(q2);q2)−q2,qG=j2(ϕM2(q2);q2). \right.
$
|
We will now do the same for the coupling at
To obtain these regions, we define the following functions with respect to conditions (FⅠ)-(FⅡ):
$hI(q2):=j2(1;q2)−j2(ϕM2(q2);q2)=q2−j2(ϕM2(q2);q2),hII(q2,q3):=j2(ϕM2(q2);q2)−j3(ϕ3M(q3);q3). $
|
These functions are continuously differentiable by Lemma 2.1. The following lemma gives the qualitative properties of the boundaries of the regions.
Lemma 3.1. There exists a unique
$ {\rm{(FI)}}\;\;\Leftrightarrow\;\;{h_{\rm{I}}}(q_2)\geqq0\;\;\Leftrightarrow\;\;q_2\geqq\tilde{q}, $ | (3.6) |
$ {\rm{(FII)}}\;\;\Leftrightarrow\;\;{h_{\rm{II}}}(q_2, q_3)\geq0\;\;\Leftrightarrow\;\;q_3\leq {\tilde h}_{\rm{II}}(q_2), $ | (3.7) |
$ {\rm{(FIII)}}\;\;\Leftrightarrow\;\;q_2\geq j_3(\phi_{\rm{3M}}(q_3);q_3), $ | (3.8) |
$
{\tilde h}_{\rm{II}}(q_2) = {0ifq2≤qneg,increasingifqneg<q2<ˉˉq,q2ifq2≥ˉˉq,
$
|
$
j_3(\phi_{\rm{3M}}(q_3);q_3) = {q3ifq3≤ˉq,increasingifˉq<q3<ˉˉq,jg(ϕinfl)+q3ϕinflifq3≥ˉˉq.
$
|
Proof. Using Lemma 2.1, we get
$
hI′(q2)=1−ϕM2>0,hI(0)=−j2(ϕM2(0);0)=−jg(ϕM2(0))<0,hI(ˉˉq)=j2(1;ˉˉq)−j2(ϕinfl;ˉˉq)>0,
$
|
where the last inequality follows from the fact that
$∂hII∂q2=ϕM2(q2)≧0⇔q2≧qneg,∂hII∂q3=−ϕ3M(q3)<0for all q3. $
|
Since the latter derivative is always non-zero, the implicit function theorem implies the existence of a continuously differentiable function
$hII(q2;˜hII(q2))=0for all q2,˜h′II(q2)=−∂hII/∂q2∂hII/∂q3≧0⇔q2≧qneg. $
|
For
$ {h_{\rm{II}}}(q_2, 0) = j_2(0;q_2)-j_3(1;0) = 0-0 = 0\;\;\Rightarrow\;\;{\tilde h}_{\rm{II}}(q_2) = 0, $ |
and for
$ {h_{\rm{II}}}(q_2, q_2) = j_2(\phi_{\rm infl};q_2)-j_3(\phi_{\rm infl};q_2) = \phi_{\rm infl}(q_2-q_2) = 0\;\;\Rightarrow\;\;{\tilde h}_{\rm{II}}(q_2) = q_2, $ |
Condition (FⅢ) can directly be rewritten by using the identity
$j3(ϕM3(˜q);˜q)−˜q=j2(ϕM2(˜q);˜q)−˜q=hI(˜q)=0. $
|
As we did for the gas injection point, we now transform the boundary curves of conditions (FⅠ)-(FⅢ) from the
$
\left\{qG=qSSU+q2,qF=q3−q2. \right.\;\;\Leftrightarrow\;\;
\left\{qG−qSSU=q2,qF=q3−q2. \right.
$
|
This mapping depends on the chosen steady-state value of
For the coupling at
$
\left\{qF=qSSU−qSSG+q3,qW=q4−q3, \right.\;\;\Leftrightarrow\;\;
\left\{qF−(qSSU−qSSG)=q3,qW=q4−q3, \right.
$
|
the chart in the
It seems logical that when working with steady states, it is necessary to restrict the amount of gas, fluid or water pumped into the tank. Moreover, from the desliming point of view, situations as huge quantities of fluid leaving at the top of the column, loosening the froth or washing it in excess, or gas bubbles flowing down and out of the tank through the bottom tailings underflow, are not convenient. Conditions (G), (FⅠ)-(FⅢ) and (WⅠ)-(WⅡ) set a theoretical upper limit for the values of
Whereas the operating charts in Section 3.7 give an overview on how to choose the bulk velocities with respect to conditions for certain steady-state couplings at the points of injection and outlets, we here collect in Tables 1 and 3 all possible steady-state combinations between these couplings for the cases
0 (G) | |||||
0 (G) | |||||
0 |
1 | ||||
1 | |||||
0 (G) | |||||
0 (G) | |||||
The tables should be read as follows. A possible steady-state solution with constant concentration in each zone is obtained by connecting adjacent rectangles passing only over vertical lines (and no corners), from the left column
Table 3 shows the possible steady states when
Despite the diversity of possible steady states with a constant value in each zone, which are categorized in Tables 1 and 3, there exist further steady states with possible stationary discontinuities within the zones. For example, in zone 1, the constant solutions
In the same way, stationary discontinuities from a lower to a higher value are possible in zones 2, 3 and 4 for all bulk velocities
For the discretization of the model, we follow the procedure of [3] for the sedimentation process in a clarifier-thickener.
We subdivide the tank into
$ϕi(t):=1Δz∫zizi−1ϕ(z,t)dz. $
|
For each layer
$
dϕidt=−F(zi,t,ϕi)−F(zi−1,t,ϕi−1)Δz+1Δz∑S∈{G,F,W}∫zizi−1qSϕSδ(z−zS)dz,
$
|
(4.1) |
where the flux
$
Fi:=Fi(ϕi,ϕi+1,t)={minϕi≤ϕ≤ϕi+1jk(ϕ)ifϕi≤ϕi+1,maxϕi≥ϕ≥ϕi+1jk(ϕ)ifϕi>ϕi+1,
$
|
(4.2) |
where
Substituting the numerical fluxes into the exact version of the conservation law (4.1), we get the following method-of-lines formula:
$
dϕidt=−Fi−Fi−1Δz+∑S∈{W,F,G}qSϕSΔzδi,iS,
$
|
(4.3) |
where
Using an explicit Euler step for the approximation of the time derivative and the CFL condition
$ \Delta t \max\limits_{k\in\{{\rm{U}}, 1, 2, 3, 4, {\rm{E}}\}}\left(\max\limits_{0\leq\phi\leq 1}|j_k'(\phi)|+|q_k|\right)\leq\Delta z, $ | (4.4) |
we obtain the following fully discrete method for
$ϕi,n+1=ϕi,n+ΔtΔz[−Fni+Fni−1+∑S∈{W,F,G}qnSϕnSδi,iS],i=−1,…,N+2. $
|
The examples demonstrate the dynamic and steady-state behaviour of a flotation column and we use dimensionless numbers. The drift-flux function
First, we define an initial set of velocities for the inlets and the outlets. In order to maximize the number of possible steady states, and also satisfying conditions (G), (FⅡ), (FⅢ), (WⅠ) and (WⅡ), we use the operating charts in Figures 10-12 to first choose
$q(z,t)={q1=−0.1forz<zG,q2=0.1forzG≤z<zF,q3=0.2forzF≤z<zW,q4=0.2353forz≥zW, $
|
and hence
![]() |
In addition to the volume fractions
Let us now run two simulations.
Example 1. Initially, we consider a column filled with only fluid, i.e.,
Example 2. Figures 18 and 19 show the results when we close the top of the tank for a longer period, in this case until
We now define an initial set of velocities for the inlets and the outlets for which no steady state with
$q(z,t)={q1=−0.1forz<zG,q2=0.1forzG≤z<zF,q3=0.15forzF≤z<zW,q4=0.17forz≥zW, $
|
and hence
As in Examples 1 and 2, we start with a column filled with only fluid, i.e.,
A flotation column is generally operated so that there are two regions: one with bubbles (intermediate gas volume fraction) and one with froth. In the bubbly region, usually in zone 2 and sometimes the lower part of zone 3, the hydrophobic particles of the pulp slurry attach to the air bubbles. In the froth region, located above the bubbly region, further enrichment takes place and the foam is efficient for promoting water rejection. The injection of wash water into the foam assists with rejection of entrained slimes. This is, however, effective only if the wash water flows downwards through the foam and bubbly regions, as stated by Dickinson and Galvin in [13]. Then the the foam is washed properly. In all steady states of Experiment 1, the wash water flows upwards.
To achieve a proper wash, the flux of water in zone 3, and consequently in zones 1 and 2, should be negative, while in zone 4, it should remain relatively small. We set
$q(z,t)={q1=−0.11forz<zG,q2=−0.01forzG≤z<zF,q3=0.04forzF≤z<zW,q4=0.1114forz≥zW, $
|
hence
The present work has shown how the available drift-flux theory for flotation columns, to the authors' knowledge so far utilized in the engineering literature for stationary analyses only, can be combined with results coming from the mathematical and numerical analysis of conservation laws with discontinuous flux to obtain a model for transient simulations as well as prediction of steady states.
The well-posedness established in [10] for the problem of continuous sedimentation, which has one flux discontinuity, covers the case here with flotation with several (but finite number of) discontinuities. In [10], a Kružkov-type of entropy condition was used together with a crossing condition for the fluxes for the proof of uniqueness. It is worth noting that the flux discontinuities within the present model do satisfy this condition. The entropy condition used here implies, however, the Kružkov-type and uniqueness is obtained without the crossing condition [18].
One-dimensional models such as the one treated herein are easier to solve than multi-dimensional multiphase flow models, and may be useful to model a flotation cell within plant-wide simulators [2]. Nevertheless, for practical use, the present model should be improved and refined in future work. Some suggested directions of future work are as follows. Starting with a property of the model in its present formulation, we recall that the present analysis is limited to drift flux functions
With respect to the hydrodynamical setup of the flotation column and its conceptual counterpart as drawn in Figure 1, we mention that Vandenberghe et al. [39] propose an interesting recirculation: according to their Figure 1, mixture is sucked from the column at a determined level, aerated, and re-injected at another position. The extraction of material at a given rate but whose composition is part of the solution gives rise to a singular sink term whose mathematical treatment is more involved than that of a singular source term (as those appearing in (1.1)). The basic difficulty is that the sink term cannot be incorporated into the flux function; rather, the sink is represented by a new non-conservative transport term (see [6]).
In several instances, our analyses invoke available mathematical and numerical results for clarifier-thickener models. If sediment compressibility is included in a clarifier-thickener model, an effect that arises if the solid-liquid suspension is flocculated, then the governing equation for such models features an additional strongly degenerating diffusion term [10] whose appropriate treatment, roughly speaking, arises from handling it as part of the convective flux. Such a term may also be motivated in the application to flotation in future works: for instance, Narsimhan [29] derives such a term to account for gradual compressibility of the foam layer that is caused by flow of liquid through a network of plateau borders due to gravitational and capillary forces [29]. On the other hand, Stevenson et al. [37] propose a convection-diffusion model for the transport of gangue in flotation froth.
Clearly, an obvious shortcoming of the present description, although it is in line with the cited treatments of literature [13,22,25,27,31,32,38,39,43,44], is that it does not explicitly model the transport and settling of solid particles. It would be highly desirable to extend the model by solids phases, for example of hydrophobic and hydrophilic particles (of minerals and gangue material), and to include their attachment to bubble and transport via the liquid and gas components. The likely outcome of such a description is a convection-diffusion-reaction system with discontinuous flux akin to a recently advanced model of continuous sedimentation with reactions [4]. On the other hand, several gas or solid phases representing size classes that segregate and form areas of different composition can be included, and lead to first-order hyperbolic systems of conservation laws with discontinuous flux, under determined circumstances (see, e.g., [11] and the references cited in that paper).
Within the present work the emphasis has been on the construction of stationary solutions. The numerical scheme utilized, Godunov's scheme with suitable modifications to handle the flux discontinuities (see Section 4), is monotone provided that
R.B. is supported by Fondecyt project 1170473; Fondef project ID15I10291; and CRHIAM, Proyecto Conicyt Fondap 15130015. In addition, R.B. and M.C.M. are supported by BASAL project PFB 03, CMM, Universidad de Chile and CI2MA, Universidad de Concepción. M.C.M. was also supported by Conicyt Fondecyt/Postdoctorado/3150140.
[1] |
A theory of ![]() |
[2] | O. A. Bascur, Example of a dynamic flotation framework, In: Centenary of Flotation Symposium, Brisbane, QLD, 6-9 June 2005, Australasian Institute of Mining and Metallurgy Publication Series, 2005, 85-91. |
[3] | A consistent modelling methodology for secondary settling tanks: A reliable numerical method. Water Sci. Technol. (2013) 68: 192-208. |
[4] | R. Bürger, S. Diehl and C. Mejías, A difference scheme for a degenerating convection-diffusion-reaction system modelling continuous sedimentation, ESAIM: Math. Model. Numer. Anal., to appear. |
[5] |
Conservation laws with discontinuous flux: A short introduction. J. Eng. Math. (2008) 60: 241-247. ![]() |
[6] |
On an extended clarifier-thickener model with singular source and sink terms. Eur. J. Appl. Math. (2006) 17: 257-292. ![]() |
[7] |
A family of numerical schemes for kinematic flows with discontinuous flux. J. Eng. Math. (2008) 60: 387-425. ![]() |
[8] |
Well-posedness in ![]() |
[9] |
Second-order schemes for conservation laws with discontinuous flux modelling clarifier-thickener units. Numer. Math. (2010) 116: 579-617. ![]() |
[10] |
A model of continuous sedimentation of flocculated suspensions in clarifier-thickener units. SIAM J. Appl. Math. (2005) 65: 882-940. ![]() |
[11] |
On some difference schemes and entropy conditions for a class of multi-species kinematic flow models with discontinuous flux. Netw. Heterog. Media (2010) 5: 461-485. ![]() |
[12] | J. M. Coulson, J. F. Richardson, J. R. Backhurst and J. H. Harker, Coulson and Richardson's Chemical Engineering. Volume 2: Particle Technology and Separation Processes, Fourth Ed., Butterworth-Heinemann, Oxford, 2000. |
[13] |
Fluidized bed desliming in fine particle flotation, Part Ⅰ. Chem. Eng. Sci. (2014) 108: 283-298. ![]() |
[14] |
On scalar conservation laws with point source and discontinuous flux function. SIAM J. Math. Anal. (1995) 26: 1425-1451. ![]() |
[15] |
A conservation law with point source and discontinuous flux function modelling continuous sedimentation. SIAM J. Appl. Math. (1996) 56: 388-419. ![]() |
[16] |
Operating charts for continuous sedimentation Ⅰ: Control of steady states. J. Eng. Math. (2001) 41: 117-144. ![]() |
[17] |
Operating charts for continuous sedimentation Ⅱ: Step responses. J. Eng. Math. (2005) 53: 139-185. ![]() |
[18] |
A uniqueness condition for nonlinear convection-diffusion equations with discontinuous coefficients. J. Hyperbolic Differential Equations (2009) 6: 127-159. ![]() |
[19] |
Numerical identification of constitutive functions in scalar nonlinear convection-diffusion equations with application to batch sedimentation. Appl. Numer. Math. (2015) 95: 154-172. ![]() |
[20] |
Fast reliable simulations of secondary settling tanks in wastewater treatment with semi-implicit time discretization. Comput. Math. Appl. (2015) 70: 459-477. ![]() |
[21] | J. A. Finch and G. S. Dobby, Column Flotation, Pergamon Press, London, 1990. |
[22] |
bed desliming in fine particle flotation Part Ⅱ: Flotation of a model feed. Chem. Eng. Sci. (2014) 108: 299-309. ![]() |
[23] | Finite difference methods for numerical computation of discontinuous solutions of equations of fluid dynamics. Mat. Sb. (1959) 47: 271-306. |
[24] |
H. Holden and N. H. Risebro, Front Tracking for Hyperbolic Conservation Laws, Second Edition, Springer Verlag, Berlin, 2015. doi: 10.1007/978-3-662-47507-2
![]() |
[25] | Liquid transport in multi-layer froths. J. Colloid Interf. Sci. (2007) 314: 207-213. |
[26] | First order quasilinear equations in several independent variables. Math. USSR-Sb. (1970) 10: 217-243. |
[27] |
The coexistence of the froth and liquid phases in a flotation column. Chem. Eng. Sci. (1992) 47: 4345-4355. ![]() |
[28] | S. Mishra, Numerical methods for conservation laws with discontinuous coefficients, Chapter 18 in R. Abgrall and C.-W. Shu (eds.), Handbook of Numerical Methods for Hyperbolic Problems: Applied and Modern Issues, North Holland, 18 (2017), 479-506. |
[29] |
Analysis of creaming and formation of foam layer in aerated liquid. J. Colloid Interface Sci. (2010) 345: 566-572. ![]() |
[30] | O. A. Oleinik, Uniqueness and stability of the generalized solution of the Cauchy problem for a quasi-linear equation Uspekhi Mat. Nauk., 14 (1959), 165-170. Amer. Math. Soc. Trans. Ser. 2, 33 (1964), 285-290. |
[31] |
Flow characterization of a flotation column. Canad. J. Chem. Eng. (1989) 67: 916-923. ![]() |
[32] |
Oil recovery from oil in water emulsions using a flotation column. Canad. J. Chem. Eng. (1990) 68: 959-967. ![]() |
[33] | Amenability testing of fine coal beneficiation using laboratory flotation column. Materials Transactions (2015) 56: 766-773. |
[34] |
Sedimentation and fluidisation: Part Ⅰ. Chemical Engineering Research and Design (1997) 75: 82-100. ![]() |
[35] |
Science and technology of dispersed two-phase systems—Ⅰ and Ⅱ. Chem. Eng. Sci. (1982) 37: 1125-1150. ![]() |
[36] |
A. Rushton, A. S. Ward and R. G. Holdich, Solid-Liquid Filtration and Separation Technology, Wiley Online Library, 2008. doi: 10.1002/9783527614974
![]() |
[37] | Convective-dispersive gangue transport in flotation froth. Chem. Eng. Sci. (2007) 62: 5736-5744. |
[38] |
On the drift-flux analysis of flotation and foam fractionation processes. Canad. J. Chem. Eng. (2008) 86: 635-642. ![]() |
[39] | Drift flux modelling for a two-phase system in a flotation column. Canad. J. Chem. Eng. (2005) 83: 169-176. |
[40] | G. B. Wallis, One-Dimensional two-Phase Flow, McGraw-Hill, New York, 1969. |
[41] | The terminal speed of single drops or bubbled in an infinite medium. Int. J. Multiphase Flow (1974) 1: 491-511. |
[42] | B. A. Wills and T. J. Napier-Munn, Wills' Mineral Processing Technology, Seventh Edition, Butterworth-Heinemann, Oxford, 2006. |
[43] |
Bubble size estimation in a bubble swarm. J. Colloid Interf. Sci. (1988) 126: 37-44. ![]() |
[44] | A gas-liquid drift-flux model for flotation columns. Minerals Eng. (1993) 6: 199-205. |
1. | Fernando Betancourt, Raimund Bürger, Stefan Diehl, Leopoldo Gutiérrez, M. Carmen Martí, Yolanda Vásquez, A Model of Froth Flotation with Drainage: Simulations and Comparison with Experiments, 2023, 13, 2075-163X, 344, 10.3390/min13030344 | |
2. | Raimund Bürger, Stefan Diehl, María del Carmen Martí, A system of conservation laws with discontinuous flux modelling flotation with sedimentation, 2019, 84, 0272-4960, 930, 10.1093/imamat/hxz021 | |
3. | Raimund Bürger, Stefan Diehl, M Carmen Martí, Yolanda Vásquez, A degenerating convection–diffusion system modelling froth flotation with drainage, 2022, 87, 0272-4960, 1151, 10.1093/imamat/hxac033 | |
4. | Raimund Bürger, Stefan Diehl, María Carmen Martí, Yolanda Vásquez, Simulation and control of dissolved air flotation and column froth flotation with simultaneous sedimentation, 2020, 81, 0273-1223, 1723, 10.2166/wst.2020.258 | |
5. | Raimund Bürger, Stefan Diehl, M. Carmen Martí, Yolanda Vásquez, A difference scheme for a triangular system of conservation laws with discontinuous flux modeling three-phase flows, 2022, 18, 1556-1801, 140, 10.3934/nhm.2023006 | |
6. | Raimund Bürger, Stefan Diehl, María del Carmen Martí, Yolanda Vásquez, Flotation with sedimentation: Steady states and numerical simulation of transient operation, 2020, 157, 08926875, 106419, 10.1016/j.mineng.2020.106419 |
0 (G) | |||||
0 (G) | |||||
0 |
1 | ||||
1 | |||||
0 (G) | |||||
0 (G) | |||||
![]() |
0 (G) | |||||
0 (G) | |||||
0 |
1 | ||||
1 | |||||
0 (G) | |||||
0 (G) | |||||
![]() |
![]() |