1 s2.0 S135943112200833X Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Applied Thermal Engineering 214 (2022) 118894

Contents lists available at ScienceDirect

Applied Thermal Engineering


journal homepage: www.elsevier.com/locate/apthermeng

Research Paper

An efficient method for computational flow-based simulation of heat


transfer in a rotary kiln with pilot scale validation
Peter J. Witt a, Julian Johnson b, M. Philip Schwarz a, *
a
CSIRO Mineral Resources, Clayton, Victoria, 3168, Australia
b
CSIRO Mineral Resources, Waterford, WA, 6152, Australia

A R T I C L E I N F O A B S T R A C T

Keywords: Rotary kilns are used to carry out heating and chemical transformation in several mineral processing operations.
Rotary kilns A series of experimental runs was carried out using the realistic but controlled arrangement of a pilot rotary kiln
Heat transfer simulation to obtain temperature data for the validation of a novel model that couples a computational fluid dynamics
computational fluid dynamics (CFD)
model of the flame to a granular bed model through heat transfer terms. The rotary kiln was 5.27 m long, with an
Residence time
internal diameter of 410 mm, and two different feeds of inert standard commercial gravel were used in the runs,
Flame temperature
with the kiln operating in counter-current operation. Each of the feeds was run through the kiln under various
operating conditions (feed rate, kiln rotation speed, fuel rate and weir height), and kiln temperature profiles were
measured during each run. The effect of various operating parameters on heat transfer within the kiln was then
determined. A fast and efficient computational model of the pilot kiln was developed to predict the burner flame
and heat transfer in the kiln; the granular bed was modelled using a one-dimensional model which was coupled
in a pragmatic way to the computational fluid dynamics gas phase model. Dependence of the predicted tem­
perature profile on operating parameters is in reasonable agreement with the measurements, and the model can
be run rapidly for efficient assessment of a large number of design and operating condition options. While there
have been an increasing number of papers published in recent years reporting very detailed models of the motion
in granular beds in rotary drums that are extremely valuable for a fundamental understanding of issues such as
segregation, such models are very computationally intensive, and are still not feasible for industrial-scale kilns in
which the particles are small. This paper indicates that good predictions of process parameters such as bed
temperature are possible without such very time-consuming computations. Moreover, the experimental tem­
perature measurements will be useful for future rotary kiln studies.

1. Introduction In general, the aim of most kiln-based processes is to raise the solids
temperature to the target temperature for the transformation (chemical
Rotary kilns are used to heat solid particulate feed material to high or metallurgical), and to maintain the temperature within an optimum
temperature in order to carry out a metallurgical or chemical trans­ range for sufficient time for the reaction to complete. The ability to
formation of the material. The solids are fed at one end of a slowly control temperature within a reasonably narrow range is therefore
rotating inclined cylindrical reactor and removed from the other end. important. Minimization of energy consumption is also important to
The particulate material moves along the kiln within a bed which lies at control costs and CO2 emissions. The effectiveness of heat transfer to the
the bottom of the kiln, being gently circulated by the rotation of the kiln solids is therefore crucial. The primary mode of heat transfer to the bed
wall. The material is usually heated by one or more flames in the free­ of solids is radiation. This can be supplemented by convection and
board space above the bed of solids. Rotary kilns are used in several conduction using lifters, chains etc., but generation and emission of dust
mineral processing operations, such as iron ore pellet induration, can prevent lifters being used in many situations. In the case of cement,
ilmenite reduction, iron ore reduction, ore calcination, as well as related suspension pre-heaters are now used, employing the heat in the exhaust
industries such as cement manufacture. Rotary dryers, operating at gas from the kiln.
lower temperatures, are also widely used. Lim et al. [1] have discussed the advantages and disadvantages of

* Corresponding author.
E-mail address: phil.schwarz@csiro.au (M.P. Schwarz).

https://doi.org/10.1016/j.applthermaleng.2022.118894
Received 21 January 2022; Received in revised form 10 June 2022; Accepted 17 June 2022
Available online 20 June 2022
1359-4311/© 2022 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
P.J. Witt et al. Applied Thermal Engineering 214 (2022) 118894

rotary kilns compared with fluidized beds - a unit operation that has increases the computational requirements even further, particularly for
displaced rotary kilns for some applications, such as alumina calcina­ industrial-scale kilns. In any case, the combustion and heat transfer as­
tion, in recent years. They point out that although rotary kilns have pects of the freeboard is best simulated separately using CFD techniques.
inferior heat transfer, they can handle solids with a wide range of par­ An alternative approach to overcome the computational requirements of
ticle sizes, including large lumps, and have more tolerance for solids that DEM simulation of the particle bed was investigated using an Eulerian
may lead to sintering or agglomeration than do fluidized beds. The (or quasi-fluid) model of the particulate bed as long ago as 1995 [11], as
poorer heat transfer can in some instances be an advantage, since it mentioned above. Other attempts to improve this Eulerian continuum
necessitates a long reactor in which the conversion can be close to ideal methodology have included the application of soils rheology [20], and
counter-current plug flow type, giving higher conversion than may be of the kinetic theory of granular flow in a two-fluid context [21,22].
possible in a fluidized bed. The authors point out, however, that the long However, it is doubtful whether the method can discriminate between
reactor configuration can lead to difficulties in controlling the reactor the behavior of different real particle types (e.g., non spherical, mixtures
temperature. of a wide range of particle sizes, etc.), given that validation has been
Temperature control and energy optimization can be assisted by based on data for large mono-sized spherical particles. Of course, the
modelling, and one dimensional (1D) models (Thornton and Batterham, ability to treat such aspects is a huge advantage for DEM [19,23]. The
[2], Gorog et al. [3], Thurlby [4], Nicholson [5]) have been used for this most serious problem in terms of routine industrial application of the
purpose for several years. Such models are still of value today, partic­ Eulerian continuum methodology, though, is that the computational run
ularly where transients need to be accounted for, such as in the case of a times for such models for thermal problems are still very large, given
solar heated rotary kiln smelting furnace [6]. A more detailed heat that the thermal timescale for industrial scale kilns is usually many
transfer model for a single transverse cross-section of a kiln, with ac­ hours, whereas transient two-fluid models are limited to very short time-
count taken of both covered and uncovered areas of the wall, was steps for stability and accuracy. This needs to be considered in the
developed by Barr et al. [7]. Boateng and Barr [8] extended the concept context that, despite such long run times, such Eulerian continuum
of a 1D model for the bed zone by coupling it to a 2D transverse model models have difficulty simulating the details of granular flows realisti­
that accounts for (transverse) bed mixing and segregation in the bed. Li cally. An alternative approach that sequentially couples CFD and DEM
et al. [9] improved the treatment of bed-wall heat transfer, and Krit­ techniques, and thus makes use of the advantages of each technique, has
zinger and Kingsley [10] showed how this type of 1D kiln model can also recently been described by Witt et al. [19]. The present paper describes a
be extended to account for multiple air jets (or flames) along the length coupled CFD-bed model that is rapid to run, and hence suitable for
of a rotary kiln, as is often used in pyrometallurgical minerals process­ assessment of a large number of kiln designs and operating conditions,
ing, e.g., in iron ore reduction. which is desirable both in the early stages of design and in process
Complex issues however require more detailed numerical modelling, optimization, as illustrated for an indirectly heated kiln application by
for example computational fluid dynamics (CFD) models, which have Proch et al. [24].
been used for the gaseous freeboard including combustion processes, Development of detailed kiln heat transfer models requires reliable
and can also be readily extended to incorporate heat transfer through data taken from kilns of a size approaching full industrial scale, a task
the shell. The first published three-dimensional (3D) CFD model of a that is not straightforward. The simplest method to obtain thermal data
rotary kiln was that developed by Bui et al. [11] for a coke calcining for operating kilns is to use an infrared scanner to measure shell tem­
rotary kiln. Four separate sub-models were used, one each for the kiln perature, as carried out by Goshayeshi and Poor [25] for a cement kiln,
bed, gas space (with reactions), kiln walls and gas space radiation. but inferring internal bed temperature is difficult in practice. Le Guen
Clinker formation in a coal fired rotary kiln was modelled by Mastorakos et al. [26] studied heat transfer in an 8 m long drum reactor for bitumen
et al. [12] using a two-dimensional axisymmetric CFD model of the aggregates in which lifters were used to maximize heat transfer (a
freeboard, in which coal combustion, radiative transfer, and heat con­ configuration common in rotary dryers). They supplemented infrared
duction were accounted for. This gas phase model was coupled to a one- shell measurements with internal gas temperature measurements using
dimensional model of the bed in which the clinker charge moved axially thermocouple probes positioned on a long rod down the center of the
at constant speed and was assumed to be well mixed (in temperature and drum, together with bitumen temperature measurements and sampling
composition) in each cross-section by the action of the kiln rotation. Bed at the walls. From these measurements they were able to calculate bulk
reactions were included subject to the 1D assumption that concentra­ heat transfer coefficients between bitumen and gas, and to the wall. The
tions are uniform at each axial cross-section. Heat transfer was also researchers used the measurements to develop and validate a 1D model
simulated through the kiln walls, including an assumed uniform- of the drum heat transfer (Le Guen et al. [27]).
thickness layer of solid clinker. In smaller kilns and drums, it is possible to carry out more detailed
Eghlimi et al. [13] performed CFD modelling of the three- and better controlled temperature measurements, particularly in the
dimensional behaviour and mixing of raw oil shale and recycled com­ case of inert solids. For example, Herz et al. [28] have measured the
busted oil shale particles in a rotary kiln for a retorting process for the temperature at multiple points within the bed of a relatively short
extraction of oil. An Eulerian-Eulerian three-phase model was used with indirectly heated rotary drum using a rod with multiple thermocouples,
one phase being gas and separate solid phases used for the two different fixed to the wall and rotating with it through the bed. The bed-wall
types of oil shale particles, represented as pseudo-fluids. Only limited contact thermal heat transfer coefficient could thus be determined
details of the model and results were presented, with no validation with good accuracy. A somewhat modified version of this monitoring
described. Amongst other published techniques for rotary kiln simula­ system has been used in a 6.7 m long pilot kiln by Liu and Specht [29]
tion is the work by Küssel et al. [14] investigating the limestone calci­ fed with inert solids (sand) and heated with a natural gas burner: ther­
nation process. A CFD model of the gas-space flame was used to provide mocouples were inserted individually into the kiln to different radial
heat transfer information to a classical one-dimensional process model distances, with this arrangement repeated at four axial locations along
of the kiln bed, in a manner analogous to the technique used by Mas­ the kiln. This enabled the radial temperature profile through the bed to
torakos et al. [12]. be monitored: for example, it was found that there was a large radial
For more detailed description of the particulate solids motion in the temperature gradient across the bed near the feed end, but the differ­
bed, Discrete Element Method (DEM) models have been used (e.g., ences had mostly dissipated at a distance of about 1.0 m from the feed.
Finnie et al. [15], Pereira et al. [16], Li et al. [17]), but the computa­ In summary, there are two main aspects to the novelty of the work
tional run time requirements are substantial, and in fact not yet feasible presented in this paper, relative to the studies that have been previously
for fine particles in an industrial scale kiln. Furthermore, heat transfer in published. While there have been models of the flame in a kiln published
the bed is more difficult to solve by these techniques (see [18,19]), and in the literature, and also some similar 1D models of the granular bed of

2
P.J. Witt et al. Applied Thermal Engineering 214 (2022) 118894

Fig. 1. Schematic diagram of pilot rotary kiln used for the measurements.

Fig. 2. (a) CFD Geometry of the burner end of the kiln, cut along the central plane. (b) Photograph of the burner.

a kiln, there are few papers that demonstrate the coupling of these two yet still capable of running rapidly so as to allow rapid assessment of a
aspects of a kiln, together with validation that clearly demonstrates that wide range of design and operating condition options, a class of model
such a coupled model can give good prediction of the main outputs that has been described as pragmatic [30]. Thirdly, measurements of
needed for process optimization of a rotary kiln. Secondly, it has been bed residence time are presented and compared with predictions from a
shown from the literature reviewed above, that there have been an semi-analytical model of bed shape. Fourthly, predictions of the coupled
increasing number of papers published in recent years reporting very computational model of the pilot kiln are compared with measurements
detailed models of the motion in granular beds in rotary drums, using of bed temperature along the kiln. The comparisons are made for various
techniques such as the Discrete Element Method (DEM), and these values of operating parameters, namely feed rate, kiln rotation speed
studies are extremely valuable for a fundamental understanding of and fuel rate, as well as some comparisons for particle size and weir
granular bed issues such as segregation. However, such models are very height. Finally, a discussion of some of the results is given, in particular,
computationally intensive, and arguably are still not even feasible for relating to energy flows within the rotary kiln. It should be emphasized
large industrial-scale kilns operating with small particles. This paper that the pilot kiln experiments were conducted solely to gather data for
aims to point out that good predictions of process parameters such as validation of the heat transfer model; the solid materials were inert were
bed temperature are possible without such very time-consuming com­ not intended to simulate any specific industrial operation, and no
putations. Thus, this paper highlights an efficient method for process attempt was made to optimize the heat transfer.
simulation of a rotary kiln.
This paper is organized in the follow way. Firstly, the experimental 2. Experimental methodology
arrangement is described for carrying out a series of pilot kiln runs at
CSIRO’s laboratories at Waterford, Western Australia, to obtain heat Temperature measurements were made in a pilot rotary kiln located
transfer data for model development and validation. Two different inert at CSIRO’s laboratories at Waterford, Western Australia. The rotary kiln
feeds materials were run through the kiln under various operating was 5.27 m long, with an internal diameter of 410 mm. The kiln angle
conditions and kiln temperature profiles were measured during each was set at 2.1◦ for all test work. A schematic diagram of the rotary kiln is
run. The effect of various operating parameters on heat transfer within shown in Fig. 1. This shows the feed end of the kiln (indicated by “12”) at
the kiln could therefore be gauged. Secondly, a computational model left, and the discharge end, at which the burner (“5”) is located, at the
methodology is described for coupling a CFD model of the freeboard right. The kiln was thus operated in counter-current mode for this work.
flame zone with a one-dimensional model for the bed. The coupled The figure also shows the approximate positions of thermocouples along
model is designed to capture the essential aspects of kiln heat transfer, the kiln as small square symbols. The thermocouples were arranged to

3
P.J. Witt et al. Applied Thermal Engineering 214 (2022) 118894

Table 1 of the steady state runs. When the beads reached the discharge end of the
Approximate particle size distributions of the two pilot kiln feeds, given as kiln, as monitored by viewing through the burner inlet, the time was
fraction of the feed in each size range. recorded as the residence time. Two differently sized sets of ceramic
Feed 1 Size range − 6.4 + 3.2 mm − 3.2 + 1.6 mm − 1.6 + 0.85 mm beads were used to ensure they were of similar size to each of the feeds,
Fraction 33.3% 33.3% 33.3% and thus gave an accurate reflection of the residence time for each of the
Feed 2 Size range − 1.0 + 0.5 mm gravel samples. The assumption is that an individual ceramic ball of the
Fraction 100% same size and density as the feed would be carried along by the feed.
Lebas et al [31] measured the residence time using both the coloured
particle technique and by weighing the kiln contents and dividing by
protrude slightly through the kiln wall into the interior of the kiln. The
feed-rate and found that the two measurements agreed well. It should be
bed temperature was measured at nine thermocouples along the length
emphasized that in the present work, residence time was measured
of the kiln to allow the effect of various operating parameters on heat
mainly as a further characterization of the operating condition, and not
transfer within the kiln to be determined. These thermocouples were
as a central part of the investigation described in this paper. Nonethe­
labelled sequentially from 1 to 9, where thermocouple 9 is closest to the
less, the measured values of residence time are discussed later in this
burner. A photograph of the natural gas burner used is shown in Fig. 2.
paper, and the usefulness of residence time as an adjunct to the CFD
Further details of the burner are provided later in this paper when the
model is indicated.
CFD geometry is described.
Two different feeds, both consisting of commercial gravel (predom­
3. Numerical modelling methodology
inantly silica) were used in the runs, with the samples differentiated by
their particle size distribution. Each of the feeds was run through the kiln
A CFD model of the pilot rotary kiln, developed to predict the burner
under various operating conditions, and kiln temperature profiles were
flame and the heat transfer from the flame, is described in this section.
measured during each run. The two feed materials were considered inert
This CFD model of the gas space (the freeboard) is coupled thermally to a
at the operating temperatures used. The first feed sample was sized in
simple 1D model of the particulate bed which accounts for the solids-
the range –6.4 + 0.85 mm, and the second, –1 + 0.5 mm. (The standard
mediated transport along the kiln. This model structure reflects the
mineral processing notation is used here to indicate, for example, –6.4 +
fact that the particulate bed and the gas space above are almost entirely
0.85 mm means the size ranges from 0.85 mm to 6.4 mm.) The coarser
decoupled in terms of their dynamics and movement. The bed moves on
feed had a wider distribution of particle size than the finer feed. The
the bottom of the kiln under the combined influence of rotation and
approximate particle size distributions of the feeds are summarized in
gravity, uninfluenced by the gas motion above. The sole coupling in this
Table 1.
model structure is thermal (including radiation), though for reactive
There are no lifters in the kiln, and rotation rate and particle size are
solids, there could also be mass transfer resulting from generation of
such that particles are not entrained into the gas space above the bed
gases in the bed, and such a situation can be easily incorporated into the
(the freeboard). Consequently, the particulate bed moves along the
model structure.
bottom of the kiln, dynamically unaffected by the gas flow. The bed is
CFD has been used widely to simulate complex heat transfer prob­
shallow in all cases, as is the usual practice in most large industrial ro­
lems in the past decade. As an example, Tian et al. [32] simulated
tary kiln operations. The bed depth for the standard base is only about
combustion of pulverized brown coal in a furnace in which heat transfer
20 mm over most of the kiln length (as will be shown through pre­
between the gas phase, the coal particles and the walls takes place
dictions given in the Appendix) compared with a diameter of 410 mm.
through turbulent convective, radiative and conductive transfer. In
The feed rate, kiln rotation speed (rpm) and fuel rate were varied
another case, heat transfer in the molten pool in laser welding [33] is
during the tests to determine their effect on the temperature profile. In
simulated with consideration of solidification, evaporation, radiation
all, 13 tests were undertaken and the operating conditions for each of
and convection. The present model for a rotary kiln also involves
these tests are given in Table 2. Note that the parameter used to char­
considerable complexity, in this case of granular flow, radiation, com­
acterize burner condition is ‘operating temperature’, taken to be the
bustion and convection, but the aim is somewhat different: the goal is to
approximate temperature between thermocouples 6 and 8. That is, for
develop a fast, efficient yet reasonably accurate simulation method.
all runs, the burner natural gas flowrate is varied to obtain the required
target ‘operating temperature’.
Residence times were monitored during all tests, and the values are 3.1. Gas phase freeboard equations
summarized in Table 2. The times were measured using a small number
of ceramic beads that were placed into the star feeder once during each The average speed of the gas flow along the kiln can be shown to be

Table 2
Summary of operating conditions for pilot kiln experimental runs.
Run # Feed Size (mm) Feed Rate (kg/hr) Weir Height (mm) Temp (◦ C)* Rotation Speed (rpm) Natural gas flow (m3/hr) Air flow (m3/hr) Residence
Time
(min)

1 − 1 + 0.5 25 50 600 1.41 6.74 40.94 45


2 − 1 + 0.5 25 50 800 1.41 7.93 46.12 45
3 − 1 + 0.5 25 50 800 1.94 7.41 46.15 25
4 − 1 + 0.5 25 50 800 0.73 7.75 46.23 80
5 − 6.4 + 0.85 25 50 600 1.42 5.61 27.05 35
6 − 6.4 + 0.85 25 50 800 1.41 6.11 47.07 35
7 − 6.4 + 0.85 25 50 800 1.92 5.94 46.84 20
8 − 6.4 + 0.85 25 50 800 0.74 6.45 46.74 65
9 − 6.4 + 0.85 10 50 800 1.39 6.02 46.72 35
10 − 6.4 + 0.85 40 50 800 1.41 6.87 46.75 35
11 − 6.4 + 0.85 25 50 1030 1.40 6.85 46.70 35
12 − 6.4 + 0.85 25 100 800 1.41 6.71 46.41 75
13 − 6.4 + 0.85 25 0 800 1.41 6.96 46.17 25
*
Approximate temperature between thermocouples 6 and 8.

4
P.J. Witt et al. Applied Thermal Engineering 214 (2022) 118894

of order 2 m/s, so the Reynolds number is of order 25,000. The gas flow Magnussen and Hjertager [40] was used by Alyaser [41] and good
is therefore fully turbulent, so, to establish an efficient model of the agreement was obtained with gas phase temperature measurements. In
rotary kiln, it is simulated here with the well-established Reynolds the EDM, the rate of reaction is determined by the rate of turbulent
averaged technique, whereby the averaged velocities and turbulence mixing/dissipation, namely k/ε. The effective reaction rate, for the
quantities are solved. Consequently, the gas-phase freeboard of the kiln combined model, is computed to be the minimum of the FiniteChemistry
is modelled using the standard steady-state continuity, Navier-Stokes Rate and the Eddy Dissipation rate. The Ansys-CFX in-built reaction ki­
(momentum) and energy equations in their Reynolds Averaged (RA) netics constants (pre-exponential factor and activation energy) for
form [35,36,37]: methane combustion are used in the finite chemistry rate model.
Although the EDM was initially proposed for unmixed turbulent com­
∇⋅(ρu) = 0 (1)
bustion, Magnussen and Hjertager [40] argued that it should be equally
applicable in most cases for pre-mixed turbulent combustion. In fact,
∇⋅(ρuu) = − ∇p + ∇⋅μ∇u − ∇⋅(ρu’¯u’ ) (2)
Klarmann [42] has compared various models with qualitative data for a
partially pre-mixed flame, and the Combined Eddy Dissipation / Finite
∇⋅(ρuh) = ∇⋅λ∇T − ∇⋅ρu’¯h’ + Srad + Sreact (3)
Rate Chemistry Model performed reasonably well in terms of predictions
where u is the gas velocity vector averaged over turbulent fluctuations, of heat release distribution.
The combustion model requires the solution of scalar transport
u’ , ρ is gas density, p is the pressure, μ the gas laminar viscosity, h the gas
equations additional to the heat equation, namely equations for the
enthalpy, T gas temperature and λ gas thermal conductivity. Energy
transport of species concentrations, the primary ones being methane,
transfer through radiation is included via the enthalpy source term, Srad ,
oxygen, water vapour and carbon dioxide; nitrogen concentration can be
while the heat of combustion is included through the source term Sreact .
obtained by difference. The species transport equations are.
Given the high speed of the burner jets, the possibility of natural con­
( )
vection is neglected, i.e., the effect of gravity is not included in the gas ∇⋅(ρuYi ) = ∇⋅ Γi,eff ∇Yi + Si (9)
phase momentum equations. The equations are solved using the general
flow solver using CFX4.4 [34]. Gas density is strongly dependent on the where Yi is the mass fraction of species i. The source term Si is due to the
gas chemical composition and temperature and is calculated in each cell chemical reaction rate involving component i, and Γi,eff is the effective
using the perfect gas law. The solver uses ANSYS databases to calculate diffusivity of species i,
gas thermal conductivity and viscosity at each point in the freeboard
μT
according to the local temperature and gas chemical composition. Γi,eff = Γi + (10)
ScT
The k-ε turbulence model [38] is used to calculate the Reynolds stress
terms on the right-hand side of eqn. (2) and the Reynolds flux term on where ScT is the turbulent Schmidt number and Γi is the molecular
the right-hand side of eqn. (3). The Reynolds stress terms are approxi­ diffusion coefficient Γi which is usually much smaller than the second
mated by [38]: term in eqn(10) when the flow is turbulent.
2 The eddy dissipation reaction rate of reaction r is given by.
− ρu’¯u’ = − μT ∇u − ρkδ (4) ( )
3 ∊ [I]
Rr = A min (11)
where δ is the Kronecker delta matrix, and where μT is the turbulent or k I νrI
eddy viscosity obtained from [38]:
where [I] is the molar concentration of component I, and I only includes
μT = Cμ ρk2 /∊ (5) the reactant components, and where νrI is the stoichiometric coefficient
for component I in the reaction r. As mentioned above, the effective
Values for the turbulent kinetic energy, k, and turbulent dissipation
reaction rate, for the combined model, is computed to be the minimum
rate, ∊, are obtained by solving the following transport equations [38]:
of the Finite
[( ) ]
μ Chemistry Rate and the Eddy Dissipation rate. The Ansys-CFX in-
∇ • (ρuk) = ∇ • μ + T ∇k + Pk − ρ∊ (6) built reaction kinetics constants (pre-exponential factor and activation
σk
energy) for methane combustion are used in the finite chemistry rate
[( ) ]
μT ∊ model.
∇ • (ρuε) = ∇ • μ+ ∇ε + (C1 Pk − C2 ρε) (7)
σ∊ k
3.2. Gas phase boundary conditions
where C1 , C2 , σ k and σ ∊ are turbulence model constants, default values
being 1.44, 1.92, 1.0 and 1.3 respectively. Pk is turbulence production The inlets at the burner are modelled using CFX “Inlet” conditions
due to viscous production. specifying mass flow, together with convected values of concentration,
Similarly, the Reynolds flux term in equation (3) is approximated by enthalpy, turbulence kinetic energy and dissipation rate of turbulence
[36]: energy. Burner inlet air temperature was 30 ◦ C with the fuel and air flow
μT C p rate given in Table 2. Regions of tramp air inflow (or air in-leakage) are
− ρu’¯h’ = ∇T (8) treated in the same way but at an ambient temperature of 20 ◦ C. The gas
σH
outlet is treated as a CFX “Outlet” at which pressure is specified, with
where σH is the turbulent Prandtl number, taken to be the conventional zero diffusion flux on convected variables. At the solid walls, standard
value of 0.85. Once again, the solver uses ANSYS databases to calculate turbulence wall functions are used for velocity, enthalpy, turbulence
specific heat capacity, Cp , according to the local temperature and gas kinetic energy and dissipation rate of turbulence energy.
chemical composition. Values of velocity were initialized to zero and the values of tem­
The discrete transfer model (a ray-tracing technique) [35,39] is used perature and turbulence were initialized to the inflow values. The
to calculate transport of thermal energy by radiation, and the Combined combustion scalars were initialized assuming the kiln was initially filled
Eddy Dissipation/Finite Rate Chemistry Model is used for pre-mixed with air. A steady state model is used and while initial conditions are set,
combustion calculations [34]. The EDM (eddy dissipation model) of they have no effect on the results but can affect convergence of the
model.

5
P.J. Witt et al. Applied Thermal Engineering 214 (2022) 118894

volume method on a co-located grid. A segregated iterative procedure is


used to solve the individual transport equations (eqn. (1), (2), (3), (6),
(7) & (9) where for each iteration the velocity equations are solved,
followed by a pressure correction step where a single equation is derived
using the SIMPLEC algorithm that adjusts the velocities to satisfy mass
conservation. Once mass conservation is enforced, the remaining scalar
equations and the radiation transport model are solved. This procedure
is repeated until residuals for all equations are reduced to suitably small
values. To capture gradients the higher order van-Leer difference
scheme was used for convective terms in the equations.

3.5. Heat transfer in the granular bed

Fig. 3a. Effect on residence time of change in rotation speed and particle size A heat transfer model of the wall boundary and bed was developed
(800 ◦ C, 25 kg/hr feed rate, 50 mm weir). by incorporating a longitudinal one-dimensional model for the bed into
the CFD model. The 1D model acts effectively as a rather complex
3.3. Gas phase geometry and grid temperature boundary condition for the CFD model and is solved
simultaneously with the CFD model. A critical assumption implicit is the
The geometry of most of the kiln is circular cylindrical. The geometry 1D model is that the particulate bed is well mixed in the transverse di­
of the burner end of the kiln is shown in Figure 3. The use of a three- rection, which appears, from the measurements of Liu and Specht [29],
dimensional 45◦ wedge segment (rather than a two-dimensional axi- to be valid except near the feed. A second assumption is that the tem­
symmetric model) allows the 3D effect of the eight individual holes in peratures of the bed and the inside boundary of the wall are closely
the burner nozzle to be captured, while minimizing the mesh size and coupled, a result found in instrumented experiments by Barr et al. [43],
computational run time. The 180◦ section, shown in these figures to even in the presence of bed reactions. Similar 1D bed models have been
improve clarity, was generated by rotating the 45◦ wedge four times. used many times by previous researchers, e.g., Thornton and Batterham
Implicit in the choice of this grid are assumptions that: (1) the particu­ [3], Thurlby [5], Nicholson [6], Bui et al. [11], Goshayeshi and Poor
late bed takes up a small fraction of the volume of the kiln, i.e., that the [25].
“fill” is small, which is indeed the case; and (2) that the internal tem­ The model is based on the transport of bed mass and energy from the
perature of the kiln wall is the same as the temperature of the particulate feed end to the burner end of the kiln. Using the CFD model mesh as a
bed. ANSYS meshing software was used to generate a mesh of approx­ guide, the kiln bed is divided along the axis into 194 one-dimensional
imately 192,000 hexahedral cells. Cells on the centerline were collapsed slices. Each slice receives mass and thermal energy (as sensible heat)
by the solver into prisms as part of the mapping to cylindrical co­ from the upstream slice. Mass in each slice is assumed to be constant, but
ordinates. A rectangular blocking structure was used to capture the kiln this assumption could be easily modified to account for bed reactions.
and burner geometry and to optimize cell shape. This also allowed The bed and inside refractory brick surface temperatures are assumed to
concentration of the mesh in the area of the burner to resolve the burner be equal as a result of the continual contact. Thus, the heat balance at
holes and where gradients were the highest. The rectangular blocks were each slice can be written:
then subdivided into cells to give approximately 195 cells along the kiln
ΔQi = ΔQsol sol rad con
i− 1 − ΔQi + ΔQi + ΔQi − ΔQwall (12)
axis, 20 cells in the 45◦ circumferential direction and 49 cell in the radial
i

direction. Since it is a cylindrical mesh, it is orthogonal, which is an


where ΔQsol i− 1 is the heat transported with the bed solids motion from
advantage in terms of fast convergence.
slice i-1 to slice i, a quantity that can be directly calculated from the feed
Air and natural gas enter the burner through the green face in Fig. 2
rate and bed temperature. The radiative and convective sources for
(a) from an upstream mixing chamber and mass flow controllers. A
transfer to the bed from the gas, ΔQrad and ΔQcon
i , are directly obtained
photograph of the burner is shown in Fig. 2(b). As can be determined i
from the CFD solution as boundary conditions. Heat loss from the bed
from the air–fuel ratio in Table 2, insufficient air for complete com­
bustion is supplied through the burner and additional air is required. through the kiln wall, ΔQwall i , is calculated based on transfer through the
The blue face in Fig. 2(a) is an open space around the burner nozzle refractory lining and heat loss by natural convection from the external
where ambient air is drawn into the kiln, which is at a pressure slightly wall of a horizontal cylinder. The rate of change in heat energy for slice i
below atmospheric due to the ID (induced draught) fan. Because neither is zero for the steady state solution.
the amount of air entering the kiln through the area around the burner In the next section, data for bed residence time is presented and
(‘tramp’ air) nor the total flow through the kiln could be measured in compared with semi-analytic models of bed motion. It is important to
this case, airflow rate is based on the pressure in the kiln, which is emphasize that bed residence time is not needed as an input to the bed
controlled by the ID fan that draws air through the kiln. Tramp air ve­ heat transfer model described above. The heat transfer by bed motion is
locities were established for three cases to give reasonable temperature unaffected by bed residence time in a steady state, since it can be
agreement with the experiment and then used for the other runs that had determined from the feed rate and temperature alone.
similar fan speeds.
For all runs the same kiln wall boundary conditions were used, and as 4. Residence time – Data versus prediction
mentioned, only three different air velocities were used for the tramp air
with these based on changes in the ID fan speed. The only other varia­ As mentioned in Section 2, residence times were monitored during
tions in the model parameters for the different runs were changes to the all tests, and the measured values are summarized in Table 2. The
burner air and natural gas flow and the solids feed rates. measurements were made to add to the understanding of operating
conditions, but are not central to the main thrust of the paper which
addresses thermal behaviour of the kiln, as investigated in the following
3.4. CFD solution technique sections. However, as we will show later, residence time can also com­
plement the CFD model in describing the thermal behaviour of the
Solution of the conservation equations for mass, momentum, thermal material in the kiln.
energy, turbulence, and species is achieved using a pressure based finite Residence times were found to vary according to the size of the feed,

6
P.J. Witt et al. Applied Thermal Engineering 214 (2022) 118894

Scott et al. [46], the best theoretical prediction of residence time is


usually obtained using the differential equation for bed height derived
by Saeman [47]. This formulation has been applied to the present pilot
kiln, as described in the Appendix. The dynamic angle of repose for the
fine gravel is usually given in the literature as 33–37◦ (Liu et al. [48]),
though the presence of sand (which is usually present in natural gravel
as a result of attrition) reduces the angle of repose by 10–20◦ and
increasing roundness (which is likely to occur over time with transport)
also leads to lower angle of repose [49]. The predictions using this value
are rather greater than the measured residence times however, and this
could be explained by a lower angle of repose or partial slip at the kiln
wall – Descoins et al. [50] showed that better predictions are obtained
Fig. 3b. Effect on residence time of change in weir height (800 ◦ C, 25 kg/hr for a smooth wall when the angle of repose in the formulas of Saeman
feed rate, rotation speed 1.41 rpm, coarse). [47] is replaced by a wall frictional angle, and good agreement can be
achieved in the present case when this procedure is followed with an
the kiln speed, and weir height. As shown in Fig. 3a, residence time is assumed friction angle of 20◦ , as shown in Fig. 3a. This angle is similar to
found to decrease as rotation speed increases, as expected, and to be that used by Descoins et al. [50], though the friction angle was not
longer for fine feed than for coarse feed. As shown by Liu and Specht measured in the present experiments. Descoins et al. [50] report that for
[44], residence time typically varies linearly with 1/n, where n is the a smooth wall, the slipping regime can result in “surging”, or periodic
rotation speed, and Fig. 3a displays similar behavior. oscillations between adhesive and kinetic friction of the bed on the wall.
It is worth pointing out that the data in Table 2 for Runs 9, 10 and 11 It is possible such behaviour cannot be well captured by the Saeman [47]
indicate that residence time is independent of feed rate, which implies model.
hold-up increases almost linearly with feed-rate. This fact has been Within the accuracy of measurement, the residence time for the
known since at least 1980: Abouzeid and Fuerstenau [45] showed that coarse gravel (35 min) is unaffected by feed rate in Table 2 (Runs 9–11).
hold-up increases almost linearly with feed-rate over a reasonable range (Note that runs using the fine gravel were not carried out at more than
of rates. They explain this as: “To accommodate the larger feed rate, the one feed rate.) This finding is in accord with previous work which found
axial velocity component of the particles within the shear zone must that residence time is rather insensitive to feed rate (Liu and Specht
increase. For this to take place, the system needs more driving force, and [44]). As expected, a higher weir results in a longer residence time, as
this can only be created by building up a greater head difference of indicated in Table 2 (Runs 6, 12 and 13) and Fig. 3b.
material between the inlet and the outlet ends of the drum. This will lead
to increasing the total hold-up in the drum as a function of feed rate.” Of 5. Results – Bed temperature measurements and modelling
course, this can only apply for reasonably small fill ratios.
The dependence on feed particle size is most likely due to the com­ 5.1. Stability of experimental kiln runs
bined effects of different bulk densities and angles of repose. When the
coarse gravel was passed through the kiln instead of fine, the residence The results discussed in the following sections are the average
time dropped from 45 to 35 min, for ‘standard’ conditions of rotation measured kiln temperature profiles at approximate steady state condi­
rate 1.41 rpm, weir height 50 mm and feed rate 25 kg/hr. According to tions, when the desired temperature was achieved between thermo­
couples 6 and 8. In most runs, steady state was reached for the target

Fig. 4. Temperature variation with time during Run 4, steady state. (800 ◦ C, Fine Feed, 0.7 rpm, 25 kg/hr feed rate, 50 mm weir).

7
P.J. Witt et al. Applied Thermal Engineering 214 (2022) 118894

Fig. 5. Temperature variation with time during Run 8, steady state. (800 ◦ C, Coarse Feed, 0.7 rpm, 25 kg/hr feed rate, 50 mm weir).

Fig. 6. Temperature variation with time during Run 11, steady state. (1030 ◦ C, Coarse Feed, 1.4 rpm, 25 kg/hr feed rate, 50 mm weir).

temperature. In almost all cases, the kiln operated stably under the temperature plot for Run 11 (given in Fig. 6) shows a 20 – 30 ◦ C rise in
target operating conditions. Examples of the degree of stability achieved temperature for each thermocouple over the steady state period. Again,
are given in Figs. 4 and 5, which plot the time variation of temperatures no parameters were changed during this period, so it is assumed the kiln
at each of the thermocouples for Runs 4 and 8. With the exception of had not completely stabilized.
minor disturbances, the temperatures are all very stable, with variation
over the period less than 10 ◦ C.
It should be noted, however, that for two runs, namely Runs 5 and 5.2. Numerical model results
Run 11, the level of stability achieved was less ideal. For example, the
Detailed results are first given for Run 6, as a typical (base case)

8
P.J. Witt et al. Applied Thermal Engineering 214 (2022) 118894

Fig. 7. Contours of gas temperature along the kiln for Run 6.

Fig. 8. Normalised velocity vectors (with velocity magnitude indicated by color contours) in the region near the burner for Run 6. The notch indicates a distance of 1
m. from the burner end.

Fig. 9. Predicted wall temperature along the kiln length for Run 6.

example. Plots showing the detailed CFD model predictions for this run Fig. 10 also plots the predicted radiative, convective and net heat
are shown in Figs. 7 to 9. In Fig. 7, which plots contours of gas tem­ flux to the bed for run 6. (Note that the small-scale oscillatory behavior
perature along the kiln, the high temperature (red) region extending of the heat transfer is due to the mesh for the radiation model being
from the burner along the kiln centerline indicates the flame. The coarser than the fluid mesh and is thus a computational artefact.) The
product gases from the flame then mix with the ambient air to give a major heat transfer mechanism to the bed and wall is radiation, with
reasonably uniform gas temperature through the last two-thirds of the convection becoming significant near the feed end, and this is true for all
kiln. Normalized velocity vectors plotted in Fig. 8 show the flow field runs. At the burner end, convective heat transfer is also significant, but
and the presence of a recirculation region between the tramp-air inlet the heat flux is negative because convection is from the bed to the air.
and the kiln wall. Kiln bed and internal refractory wall temperature This shows that convection is the main mechanism for heating tramp-air
(assumed to be equal) is plotted in Fig. 9. Results for the other cases are entrained through the annular opening around the burner. Radiation in
qualitatively similar and so have not been included in this paper. the burner area is high but the energy lost through the kiln wall and the
Fig. 10 shows the predicted and measured bed temperatures along energy required for heating the incoming air results in a drop in bed
the kiln for Run 6. Good agreement between the model and measured temperature in the last meter of the kiln before the solids are discharged.
temperatures is obtained for this run, and indeed for most runs, as will
be seen in other figures presented later in the paper. The predicted
5.3. Results – Effect of operating parameters and comparison with data
external kiln wall temperature is also shown in Fig. 10: it is approxi­
mately 130–150 ◦ C for Run 6. For other runs, external kiln temperature
5.3.1. Operating temperature (Fuel rate)
ranges between 123 ◦ C (Run 5) and 187 ◦ C (Run 11), with an average for
The effect of operating temperature can be shown by comparing the
all the runs of 153 ◦ C. This is in good agreement with the estimated shell
results for 600, 800 and 1100 ◦ C (Runs 5, 6 and 11 respectively), carried
temperature of 150 ◦ C made during the experiments.
out using the larger (–6.4 + 0.85 mm) particles. Fig. 11 shows the

9
P.J. Witt et al. Applied Thermal Engineering 214 (2022) 118894

Fig. 10. Predicted bed and kiln external wall temperature and the radiative, convective and net heat flux to the bed surface for Run 6.

Fig. 11. Effect of operating temperature on the kiln bed temperature profile for –6.4 + 0.85 mm particles.

Fig. 12. Effect of operating temperature on the kiln energy flows for –6.4 + 0.85 mm particles.

measured and predicted bed temperatures for the three operating tem­ burner ends.
peratures. Good agreement for the 600 and 800 ◦ C cases is evident. For As in all cases studied, a peak in temperature is seen at around 3.8 to
1100 ◦ C, the model tends to over-predict the temperature between 1 and 4.2 m from the feed end. It is expected that this point marks the most
3 m from the feed entry, but shows good agreement at the feed and intense part of the flame, which (as can be seen from Fig. 8) extends

10
P.J. Witt et al. Applied Thermal Engineering 214 (2022) 118894

Fig. 13. Effect of feed rate on the temperature kiln profile for –6.4 + 0.85 mm particles.

Fig. 14. Effect of feed rate on the kiln energy flows for –6.4 + 0.85 mm particles.

about 1 m from the burner. The bed temperature increases gradually predicts a constant operating temperature as feed rate is changed, i.e., a
from the feed end, increasing more rapidly near the flame. It then drops similar degree of constancy as the measurements, which themselves
off rather rapidly as it passes the most intense part of the flame and display some variation. In this sense the model gives good agreement in
proceeds to the discharge end. There is some evidence for a flattening of all cases. In terms of the detailed comparison of individual temperatures,
the measured temperature profile around 2 m from the feed end, though there is good agreement for the 25 kg/hr run, but slight under-prediction
this is less obvious in the model results. It is possible that there is greater of temperature for 10 kg/hr case and slight over-prediction for the 40
external wall cooling in this zone not accounted for in the model. kg/hr case. This may be a result of the fact that the well-mixed uniform
The energy flows in the kiln for the three cases are shown in Fig. 12. bed temperature assumption may be become less accurate as the feed
Increase in operating temperature required an increase in the natural gas rate (and thus the bed depth) increases. As in most of the model pre­
flow rate and hence an increase in the energy from the burner, and this is dictions, the model somewhat over-predicts the temperature about
seen in Fig. 12 as resulting in an increase in the energy loss through the halfway along the kiln (at about 2.3 m from the feed position). Ironi­
kiln wall and energy leaving with the solids. Energy in the outlet gas cally, if the natural gas rate had not been modified, the curves for
stream reduced slightly due to the reduction in induced airflow because different feed rates would have been further apart, and it would have
the ID fan speed was reduced. appeared that the model gas much better agreement with the data. The
other aspect of Fig. 13 that may require explanation is that it is clear that
5.3.2. Solids feed rate (fuel rate compensated) for feed rate 40 kg/hr, the natural gas rate was actually over compen­
The effect of solids feed rate can be seen by comparing Runs 9, 6 and sated (i.e., increased too much) so that the bed temperature for that case
10, corresponding to solids feed rate of 10, 25 and 40 kg/hr. Note that in is greater than for the lower feed rates. However, both model and
order to maintain a constant ‘operating temperature’ (the approximate experiment show the greater temperature, again giving confidence in
temperature between thermocouples 6 and 8) of 800 ◦ C, and to avoid the the model predictions.
bed temperature shifting substantially out of the normal operating Energy flows plotted in Fig. 14 show that the higher feed rate case
range, burner natural gas rate was increased from 6.02 to 6.11 to 6.87 required an increase in burner energy with a slight increase in the energy
m3/hr as the feed rate increased from 10 to 25 to 40 kg/hr, respectively, lost through the kiln wall. Energy leaving with the solids increased with
as shown in Table 2. The three cases thus still test the ability of the model the increased solid mass flow rate, as expected, since bed temperature
to account for the dependence on feed rate. The measured and predicted was maintained at a roughly constant value. Induced airflow through
temperature profiles for these cases are shown in Fig. 13. Obviously, if the kiln was constant for each run and the energy leaving with the outlet
the burner natural gas rate was not decreased as feed rate decreased, the gas was reasonably constant.
bed temperature would rise substantially. The appropriate test of the
CFD model for the present set of experiments, therefore, is whether it

11
P.J. Witt et al. Applied Thermal Engineering 214 (2022) 118894

Fig. 15. Effect of rotation rate on the temperature kiln profile for –1 + 0.5 mm particles.

Fig. 16. Effect of rotation rate on the temperature kiln profile for –6.4 + 0.85 mm particles.

Fig. 17. Effect of particle size on the temperature kiln profile at 600 ◦ C operating temperature.

5.3.3. Rotation rate have a large effect on measured bed temperature, when the feed rate and
Kiln rotation rate was varied from 0.7 to 1.3 to 2.0 rpm for both the fuel rate are kept constant. This may seem counter-intuitive, but the
small and large particles. Predicted and measured temperatures for the reason is that if the bed is sufficiently well mixed to have uniform
three rotation speeds are compared for the two particle distributions in temperature, the heat convected along the bed will be essentially in­
Figs. 15 and 16 respectively. It is clear that the rotation speed does not dependent of the rotation speed, since it will depend mainly on solids

12
P.J. Witt et al. Applied Thermal Engineering 214 (2022) 118894

Fig. 18. Effect of particle size on the temperature kiln profile at 800 ◦ C operating temperature.

Fig. 19. Experimental effect of weir height on the kiln bed temperature profile at 800 ◦ C operating temperature.

Fig. 20. Fraction of burner energy distributed into the outlet gas, solids and lost through the kiln walls for different operating conditions.

feed rate and bed temperature. Of course, in an actual processing the CFD-based model lumps the bed and hence correctly accounts for the
operation, time-at-temperature is usually a critical parameter, and as the heat convected by the bed. The model does not account explicitly for
rotation speed increases, the residence time decreases (because the bed second-order effects of rotation speed such as bed mixing, but the
depth decreases), so the time-at-temperature seen by the bed material experimentally measured bed temperatures show that these effects are
also decreases. small. The small differences between the CFD runs shown in Fig. 15 arise
The bed temperature predictions of the CFD-based model agree with from the slight differences in experimental fuel rate (e.g., Run 2 has the
the experiments in that they are independent of rotation speed, because highest fuel rate, hence the model predicts the highest temperature). It is

13
P.J. Witt et al. Applied Thermal Engineering 214 (2022) 118894

and the heat transfer appears to be only slightly higher for the finer feed
material.
Model results for the large particles are in reasonable agreement with
the measured temperatures. For the smaller particles, agreement is good
in the burner region but the CFD model over-predicts the temperature in
the region of 1 to 3 m from the feed end. As discussed previously, this
may be a consequence of Runs 1 and 2 being the first runs and that the
kiln had not reached thermal equilibrium.
It should be noted that, following a similar reasoning as used in
relation to rotation speed, particle size will have a significant effect on
bed depth (as reflected in the measured residence times in Table 2), but
bed temperature is not expected to depend greatly on particle size
provided the bed is well mixed and hence at relatively uniform tem­
perature. The model, being integrated over bed depth, is based on this
assumption. The model dependence shown in Figs. 17 and 18 therefore
reflects natural gas rate, which was varied in the pilot runs in order to
target the desired operating temperature.

5.3.5. Weir height


Fig. A1. Schematic diagram indicating the definition of bed depth, h, as well as Fig. 19 shows the variation in the solids temperature profile along
the position of the bed at the discharge dam. the kiln as the weir height is increased from 0 to 50 to 100 mm for coarse
feed and at 800 ◦ C operating temperature. The measured temperature
profile shape and temperature itself does not change with change in feed
rate. This might seem remarkable given that the residence time changes
from 25 min at 0 weir height, to 35 min at 50 mm height, and to 75 min
at 100 mm height, but it should be remembered that the flowrate of
solids along the kiln remains unchanged. Increase in residence time is
associated with an increase in bed height which should not result in a
major change in radiative heat transfer from the flame to the bed, and
the heat convected along the bed will then not be affected if the bed is
well mixed. This is the assumption underlying the coupled CFD-bed
model, and indeed Fig. 19 would suggest that this assumption is a
good approximation.

6. Discussion

The results presented in the previous section show that the CFD
Fig. A2. Predicted bed depth profile in the pilot. The case is feed rate 25 kg/hr, model, coupled to a 1D model of the particle bed, is capable of rapidly
rotation speed 1.94 rpm, angle of repose 33◦ and dam height 50 mm.
predicting the major heat transfer trends in a rotary kiln. The model is
much more efficient compared with simulation methods that require
not clear whether the small changes made to burner natural gas rate in detailed two or three dimensional simulation of the granular bed, such
order to maintain constant operating temperature in the experiments as as those described in the papers [11,15–19]. The present method con­
speed is changed have any significance since the changes are not verges in 7.5 h on a single core, whereas 3D CFD simulations of both bed
monatonic with speed. and freeboard carried out by Witt et al [19] required more than 3 days.
Similar comments can be made regarding the effect of rotation speed DEM simulations of the bed in an industrial scale kiln with particles of
for large particles, as shown in Fig. 16. There is no major experimental the size used here is not even presently feasible. On the other hand, the
trend with rotation speed, and the model similarly predicts little effect present method is a more accurate approach than simpler models
because the bed part of the model is integrated over bed depth. Again, described in the papers [2,7–9]. The model can therefore be used as a
the slight variation in model temperature is a reflection of the small tool to better understand heat transfer mechanisms, and to subsequently
variation in experimental fuel rate (run 8 at 0.7 rpm has a slightly higher optimize kiln operation. An example of the diagnosis that is possible
fuel rate, as seen in Table 2). It is worth mentioning that the small with the model is the breakdown of heat transfer mechanisms between
changes made to burner natural gas rate in order to maintain constant radiation and convection as shown in Fig. 10. Integration along the kiln
operating temperature in the experiments as speed is changed may be also allows global energy transfers to be analysed. The relative values of
significant: the highest rotation speed are associated with the lowest fuel energy transfers in the pilot kiln, namely the energy entering the kiln
rate. through the burner, energy leaving in the gas stream through the gas
outlet, and energy transferred to the bed and kiln walls are shown
5.3.4. Particle size graphically for all the runs in Fig. 20. On average 60% of the energy
Runs with small and large particle sizes were performed at 600 ◦ C input is used in heating the air flowing though the kiln, and 34% is lost
and 800 ◦ C to assess the effect of particle size; plots of the bed tem­ through the kiln wall, with only 6% of the energy retained in the solids
perature profile for the two temperatures are shown in Figs. 17 and 18. It leaving the kiln.
appears that for the lower temperature, feed size has a substantial effect This energy break-down is not necessarily typical of full-scale in­
on the heat transfer to the burden in the flame region: the heat transfer dustrial kiln operations. Firstly, wall heat losses are reduced greatly in
appears to be significantly higher for the fine feed compared with the an industrial kiln by using insulating refractory walls; the present pilot
coarse feed, though the substantial reduction in air rate in the coarse kiln has a steel shell. Secondly, in an actual industrial process, heat
feed case may be the cause of this difference (see Table 2). There is much transferred to the solids is normally used to carry out endothermic re­
less difference between fine and coarse feed at the higher temperature, actions such as calcination. Because of the heat sink in the solids, heat

14
P.J. Witt et al. Applied Thermal Engineering 214 (2022) 118894

can be continuously transferred from the gas to the solids bed because kiln for the purpose of creating a data-set suitable for assisting devel­
the endothermic reaction will maintain the bed at a lower temperature opment and validation of numerical models, and a computational fluid
than the gas so long as the reaction continues, provided the reaction dynamics model coupled to a one-dimensional bed model was shown to
temperature is sufficiently low. For this reason, the lower the reaction be suitable for rapidly analysing heat transfer in the kiln. In particular,
temperature, the higher the energy efficiency that can be achieved, the model is effective in predicting the influence of burner natural gas
where energy efficiency is defined as the enthalpy of the solids reaction rate on temperature profile along the kiln. Dependence of the predicted
at a specified temperature divided by the energy input. It should be temperature profile on operating parameters is in reasonable agreement
emphasized that the pilot kiln experiments were conducted solely to with the measurements, and the model can be run rapidly for efficient
gather data for validation of the heat transfer model; the solid materials assessment of a large number of design and operating condition options.
were inert were not intended to simulate any specific industrial opera­ This efficient, yet reasonably accurate, simulation method is note­
tion, and no attempt was made to optimize the heat transfer. Nonethe­ worthy because both Discrete Element Method (DEM) and Eulerian
less the energy breakdown gives an insight into likely energy efficiency continuum models of the entire rotary kiln (including bed) have sub­
issues. stantial computational run time requirements, and in fact not yet
In principle, it is usually beneficial for only a small amount of the feasible for fine particles in an industrial scale kiln. This paper has
heat loss to be associated with the solids stream, since heat in the gas can demonstrated that good predictions of process parameters such as bed
more readily be used. In the cement industry, it is standard practice to temperature are possible without very time-consuming computations,
recover heat from the off-gas in pre-calciners and pre-heaters (Madloola such as the DEM simulations which have been possible only for small-
et al. [52]). Various kiln-based processes also use off-gases for co- scale kilns with relatively large particles. The present technique is
generation. Liu et al [51] has carried out a thermal analysis of the ro­ rapid even for fine particles in very large kilns and is easily extended to
tary kiln-electric furnace for the smelting of ferronickel alloy in which reacting systems.
the kiln off-gases are used for feed pre-heating and drying. The other Both model and measurements show that increase in fuel rate causes
major loss of energy in kilns is through the shell, as indicated in Fig. 20. the expected increase in bed temperature. However, model and mea­
Heat recovery exchangers have been designed to capture this heat, surements agree in indicating that kiln rotation speed does not have a
particularly in the case of rotary kilns in the cement industry [53,54,55]. strong influence on bed temperature, within the range of speeds tested.
The CFD model, coupled to a 1D model in a pragmatic way, predicts This may seem counter-intuitive, but the reason is that if the bed is
kiln temperatures with a reasonable degree of accuracy. The 1D bed sufficiently well mixed to have uniform temperature, the heat convected
model integrates over bed depth, and this is a reliable approximation along the bed will be essentially independent of the rotation speed. Of
provided the bed is sufficiently well mixed. This approach has the course, in an actual processing operation, time-at-temperature is usually
advantage that rate of heat advection along the kiln can be determined a critical parameter, and as the rotation speed increases, the residence
from bed temperature and solids feed rate. Provided the bed is suffi­ time decreases (because the bed depth decreases), so the time-at-
ciently well mixed, it is expected that the rate of heat advection along temperature seen by the bed material also decreases. Model and mea­
the kiln is independent of rotation speed, despite the fact that the bed surements also show that weir height has little effect on bed tempera­
depth is strongly dependent on rotation speed, and this expectation is ture, for a similar reason to that applying to the rotation speed effect.
supported by the experimental measurements. The same reasoning ap­ Moreover, an analytical model for bed depth has been presented, and
plies to dependence on particle size and weir height. The validity of the shown to predict bed residence time in agreement with experimental
assumption has been verified by the experimental measurements of bed measurements. Combining the results of the coupled fluid dynamics–bed
temperature presented above, and their dependence on rotation speed, heat transfer model with the analytic prediction for bed depth and
particle size and weir height. longitudinal motion would allow the time history of particle tempera­
It is essential to recognize that rotation speed, particle size and weir ture to be determined as it moves along the kiln. In summary, it is ex­
height all influence the bed depth, and hence the bed residence time (Liu pected that the presented heat transfer model will be a useful tool in
and Specht, [44]). These parameters are therefore very important for analysing kiln issues associated with heat transfer, such as optimization
design and optimisation of processing operations, in which time-at- of temperature profile for process improvement and energy efficiency.
temperature is often critical to degree of reaction achieved by the kiln.
The coupled CFD-bed model described in this paper can be com­ Declaration of Competing Interest
plemented by the method for determining residence time described in
the Appendix using equations (A1) and (A3). This, then, allows the time- The authors declare that they have no known competing financial
at-temperature to be determined. In fact, from eqn (A1), combined with interests or personal relationships that could have appeared to influence
the CFD-bed model results, it is possible to determine the detailed time the work reported in this paper.
history of particle temperature at it passes along the kiln.
There are no doubt second order effects of fill ratio on heat transfer. Data availability
For example, a deeper bed has a larger surface area for radiative transfer
from the flame. Fill ratio will also influence bed mixing and potentially No data was used for the research described in the article.
also bed-wall heat transfer, as will particle size, but such effects are
beyond the predictive capability of the present 1D bed model, which is Acknowledgements
pragmatic in capturing major effects, but also very rapid to run. A model
of the type described by Witt et al. [19] would be required to capture the The authors thank the companies that sponsored the AMIRA
more complex effects, but at the expense of much longer computational managed project, P581: Rotary Kiln Technology, which partially sup­
run times. ported this work.

7. Conclusions

Temperature measurements were carried out in a pilot-scale rotary

15
P.J. Witt et al. Applied Thermal Engineering 214 (2022) 118894

Appendix

The model derived by Saeman [47] for flow of granular material through an inclined cylinder is used to obtain the bed profile and hence residence
time for comparison with the measurements reported in this paper. The cylinder is taken to be of radius R, with rate of rotation n, and with axis
inclined at angle β to the horizontal. The granular material has angle of repose γ, and the volumetric flow rate is Q. The depth of the granular bed is h.
Then using certain assumptions about particle motion in the bed together with geometrical considerations, Saeman [47] derived the equation:
dh 3tanγ Q tanβ
= − (A1)
dx 4π n [(2R − h)h ]3/2 cosγ

where x is the distance along the kiln from the discharge end.
Scott et al. [46] explain the assumptions used in deriving this equation, prime of which is a steady-state condition. The bed surface is assumed to be
locally planar, so that in a cross-section normal to the axis of the cylinder, the bed surface forms a straight line. As the cylinder rotates, most of the
particles move in solid body rotation with angular velocity 2πn. Thus, each particle describes a circular arc and on reaching the surface, falls down a
line of steepest descent, inclined at angle γ to the horizontal. Furthermore, it is assumed that β and dh/ dx are small. Note also that the bed depth, h, is
defined as the distance from the bed surface to the inside surface of the cylinder at which the tangent is parallel to the bed surface; h is measured along
a line normal to the cylinder axis, and normal also to the above-mentioned line in the bed surface. Fig. A1 gives a schematic diagram indicating h, as
well as the position of the bed at the discharge dam.
Eqn(A1) can then be numerically solved for the shape of the bed surface h(x) by integration. Solution requires a boundary condition at the
discharge. The condition used by Scott et al. [46] is.
h(0) = Hdam + dp (A2)

where Hdam is the height of the dam or weir, and dp is the mean particle diameter. In the absence of a dam, the value of Hdam is set to zero. Scott et al.
[46] have found good quantitative agreement between the predictions of bed depth from eqn (A1) and measurements for a variety of rotating cylinder
operating conditions with both end dams and internal dams.
Using a geometrical approximation for the differential volume of the bed (i.e., the volume in differential length dx of the kiln), and numerically
integrating for the entire length of the bed gives the total volume of the bed, which allows the residence time to be determined by dividing by the
volume feed rate:

1
t= dVbed (A3)
Q

where the integral is essentially the solids (volume) holdup in the kiln. Note that Q is the bulk volume flow rate, where the packing density is assumed
to be constant everywhere. Normally the flow rate operating condition would be controlled or measured as a mass flow rate, M, so the bulk density
within the bed is an important parameter to be estimated when applying the formulas. Various authors have attempted to derive approximate formulas
for residence time to avoid numerical integration. Saeman [47] demonstrates that approximations to the integrations gives results that agree
reasonably well with previously published measurements of residence time, while Kramers and Croockewit [56], using a somewhat different method
to approximate the integrations, find good agreement for holdup (and hence residence time) over a much wider range of operating conditions.
Likewise, the agreement obtained by Scott et al [45] between eqn (A1) and measurements of bed depth profile imply that the equation should also
predict the residence time correctly when used with eqn (A3). Direct measurements of residence time (using tracers) by Lebas et al [38] have also been
used to successfully validate the predictions of eqns (A1) and (A3).
An example of the application of eqn (A1) to determination of depth profile in the pilot kiln described in this paper is shown in Fig. A1. The case is
feed rate 25 kg/hr, rotation speed 1.94 rpm, angle of repose 33◦ and dam height 50 mm. The numerical integration procedure was validated by using it
to re-calculate the bed depth profiles computed by Scott et al. [46]. These profiles were validated against measurements of bed depth [46], so there is a
high degree of confidence in the bed depth predicted for the pilot kiln in Fig. A2.

References [9] S.-Q. Li, L.-B. Ma, W. Wan, Q. Yan, A mathematical model of heat transfer in a
rotary kiln thermo-reactor, Chem. Eng. Technol. 28 (2005) 1480–1489.
[10] H.P. Kritzinger, T.C. Kingsley, Modelling and optimization of a rotary kiln direct
[1] S.L. Lim, P. Peeler, M.P. Schwarz, Comparison of solids residence time distributions
reduction process, J. South. Afr. Inst. Min. Metall. 115 (2015) 419–424.
from fluidized bed and kiln processes, Proc. 6th World Congress of Chemical
[11] R.T. Bui, G. Simard, A. Charette, Y. Kocaefe, J. Perron, Mathematical modelling of
Engineering, Melbourne, 23–27 September 2001.
the rotary coke calcining kiln, Can. J. Chem. Eng. 73 (1995) 534–545.
[2] G.J. Thornton, R.J. Batterham, The transfer of heat in kilns, Proc. of CHEMECA 82,
[12] E. Mastorakos, A. Assias, C.D. Tsakiroglou, D.A. Goussis, V.N. Burganos, A.
National Conference Publication No. 82 (9) (1982) 260–266.
C. Payatakes, CFD predictions for cement kilns including flame modelling, heat
[3] J.P. Gorog, J.K. Brimacombe, T.N. Adams, Radiative heat transfer in rotary kilns,
transfer and clinker chemistry, Appl. Math. Modell. 23 (1999) 55–76.
Metall. Trans. B 12B (1981) 55–70.
[13] A. Eghlimi, R. Benito, C. Golab, The CFD investigation of flash dryer and rotating
[4] J.A. Thurlby, A Dynamic mathematical model of the complete grate/kiln iron-ore
kiln design, Proc. 2nd Inter. Conf. on CFD in the Minerals and Process Industries, 6–8
pellet induration process, Metall. Trans. B 19B (1988) 103–112.
December, Melbourne, Australia, 1999, pp. 455–460.
[5] T.A. Nicholson, Mathematical Modelling of the Ilmenite Reduction Process in
[14] U. Küssel, D. Abel, M. Schumacher, M. Weng, Modeling of Rotary Kilns and
Rotary Kilns, PhD Dissertation, 1995, University of Queensland.
Application to Limestone Calcination, Proc. 7th Modelica Conference, Como, Italy,
[6] S.O. Alexopoulosa, J. Dersch, M. Roeb, R. Pitz-Paal, Simulation model for the
Sep. 20–22, 2009, pp. 814–822, DOI: 10.3384/ecp09430084.
transient process behaviour of solar aluminium recycling in a rotary kiln, Appl.
[15] G.J. Finnie, N.P. Kruyt, M. Ye, C. Zeilstra, J.A.M. Kuipers, Longitudinal and
Therm. Eng. 78 (2015) 387–396.
transverse mixing in rotary kilns: a discrete element method approach, Chem. Eng.
[7] P.V. Barr, J.K. Brimacombe, A.P. Watkinson, A heat-transfer model for the rotary
Sci. 60 (2005) 4083–4091.
kiln: part II. development of the cross-section model, Metall. Trans. B 120B (1989)
[16] G.G. Pereira, M.D. Sinnott, P.W. Cleary, K. Liffman, G. Metcalfe, I.D. Sutalo,
403–419.
Insights from simulations into mechanisms for density segregation of granular
[8] A.A. Boateng, P.V. Barr, A thermal model for the rotary kiln including heat transfer
mixtures in rotating cylinders, Granular Matter 13 (2011) 53–74.
within the bed, Int. J. Heat Mass Transfer 39 (1996) 2131–2147.

16
P.J. Witt et al. Applied Thermal Engineering 214 (2022) 118894

[17] D. Li, L. Wang, Q. Wang, G. Liu, H. Lu, Q. Zhang, M. Hassan, Simulations of [35] Ansys Inc, Ansys Inc., ANSYS CFX-solver theory guide, Canonsburg, PA, 2011.
dynamic properties of particles in horizontal rotating ellipsoidal drums, Appl. [36] S.V. Patankar, Numerical heat transfer and fluid flow, Taylor & Francis, 1980.
Math. Modell. 40 (2016) 7708–7723. [37] H.K. Versteeg, W. Malalasekera, An Introduction to computational fluid dynamics,
[18] M. Kwapinska, G. Saage, E. Tsotsas, Mixing of particles in rotary drums: a The Finite Volume Method. Longman. (1995).
comparison of discrete element simulations with experimental results and [38] B.E. Launder, D.B. Spalding, The numerical computation of turbulent flows,
penetration models for thermal processes, Powder Tech. 161 (2006) 69–78. Computer Methods in Appl. Mechan. Eng. 3 (1974) 269–289.
[19] P.J. Witt, M.D. Sinnott, P.W. Cleary, M.P. Schwarz, A hierarchical simulation [39] F.C. Lockwood, N.G. Shah, A new radiation solution method for incorporation in
methodology for rotary kilns including granular flow and heat transfer, Miner. Eng. general combustion prediction procedures, Symposium (International) on
119 (2018) 244–262. Combustion 18 (1) (1981) 1405–1414.
[20] M.P. Schwarz, P. Witt, Multi-scale modelling in minerals processing, in: M. [40] B.F. Magnussen, B.H. Hjertager, in: On mathematical modeling of turbulent
P. Schwarz, P. Witt (Eds.), Proceedings of Symposium on Multi-Scale Modelling for combustion with special emphasis on soot formation and combustion, The
Industrial Flow Systems, 2009. Combustion Institute, 1977, pp. 719–729.
[21] D.A. Santos, F.O. Dadalto, R. Scatena, C.R. Duarte, M.A.S. Barrozo, [41] A. Alyaser, Fluid flow and combustion in rotary kilns, UBC, 1998. PhD thesis.
A hydrodynamic analysis of a rotating drum operating in the rolling regime, Chem. [42] N. Klarmann, Modeling turbulent combustion and CO emissions in partially-
Eng. Res. Des. 94 (2015) 204–212. premixed conditions considering flame stretch and heat loss, Technischen
[22] W. Rong, Y. Feng, M.P. Schwarz, P. Witt, B. Li, T. Song, J. Zhou, Numerical study of Universität Muenchen, 2019. PhD Dissertation.
the solid flow behavior in a rotating drum based on a multiphase CFD model [43] P.V. Barr, J.K. Brimacombe, A.P. Watkinson, A heat-transfer model for the rotary
accounting for solid frictional viscosity and wall friction, Powder Tech. 361 (2020) kiln: part I. pilot kiln trials, Metall. Trans. B 120B (1989) 391–402.
87–98. [44] X.-Y. Liu, E. Specht, Mean residence time and hold-up of solids in rotary kilns,
[23] H.R. Norouzi, R. Zarghami, N. Mostoufi, Insights into the granular flow in rotating Chem. Eng. Sci. 61 (2006) 5176–5181.
drums, Chem. Eng. Res. Des. 102 (2015) 12–25. [45] A.-Z.-M. Abouzeid, D.W. Fuerstenau, A study of the hold-up in rotary drums with
[24] F. Proch, K. Bauerbach, P. Grammenoudis, Development of an up-scalable rotary discharge end constrictions, Powder Tech. 25 (1980) 21–29.
kiln design for the pyrolysis of waste tyres, Chem. Eng. Sci. 238 (2021), 116573. [46] D.M. Scott, J.F. Davidson, S.-Y. Lim, R.J. Spurling, Flow of granular material
[25] H.R. Goshayeshi, F.K. Poor, Modeling of rotary kiln in cement industry,, energy through an inclined, rotating cylinder fitted with a dam, Powder Tech. 182 (2008)
and power Eng. 8 (2016) 23–33. 466–473.
[26] L. Le Guen, F. Huchet, J. Dumoulin, Y. Baudru, P. Tamagny, Convective heat [47] W.C. Saeman, Passage of solids through rotary kilns: factors affecting time of
transfer analysis in aggregates rotary drum reactor, Appl. Therm. Eng. 54 (2013) passage, Chem. Eng. Prog. 47 (1951) 508–514.
131–139. [48] X.-Y. Liu, E. Specht, J. Mellmann, Experimental study of the lower and upper
[27] L. Le Guen, F. Huchet, P. Tamagny, Drying and heating modelling of granular flow: angles of repose of granular materials in rotating drums, Powder Tech. 154 (2005)
application to the Mix-Asphalt processes, J. of Appl. Fluid Mech. 4 (2011) 71–80. 125–131.
[28] F. Herz, I. Mitov, E. Specht, R. Stanev, Influence of operational parameters and [49] H.M.B. Al-Hashemi, O.S.B. Al-Amoudi, A review on the angle of repose of granular
material properties on the contact heat transfer in rotary kilns, Inter. J. Heat and materials, Powder Technol. 330 (2018) 397–417.
Mass Trans. 55 (2012) 7941–7948. [50] N. Descoins, J.-L. Dirion, T. Howes, Solid transport in a pyrolysis pilot-scale rotary
[29] X.-Y. Liu, E. Specht, Temperature distribution within the moving bed of rotary kiln: preliminary results – stationary and dynamic results, Chem. Eng. and Process.
kilns: Measurement and analysis, Chem. Eng. and Process. 49 (2010) 147–150. 44 (2005) 315–321.
[30] M.P. Schwarz, Y.Q. Feng, Pragmatic CFD modelling approaches to complex [51] N.A. Madloola, R. Saidura, M.S. Hossaina, N.A. Rahim, A critical review on energy
multiphase processes. Progress in Applied CFD. (2015) J.E. Olsen and S.T. use and savings in the cement industries, Renew. Sustainable Energy Rev. 15
Johansen (Eds.) SINTEF, pp. 25–38. (2011) 2042–2060.
[31] E. Lebas, J.L. Houzelot, D. Ablitzer, F. Hanrot, Experimental study of residence [52] P. Liu, B. Li, S. Cheung, W. Wu, Material and energy flows in rotary kiln-electric
time, particle movement and bed depth profile in rotary kilns, Can. J. Chem. Eng. furnace smelting of ferronickel alloy with energy saving, Appl. Therm. Eng. 109A
73 (1995) 173–180. (2016) 542–559.
[32] Z.F. Tian, P.J. Witt, M.P. Schwarz, W. Yang, Combustion of pre-dried brown coal in [53] Z. Söğüt, Z. Oktay, H. Karakoç, Mathematical modeling of heat recovery from a
a tangentially fired furnace under different operating conditions, Energy and Fuels rotary kiln, Appl. Therm. Eng. 30 (2010) 817–825.
26 (2) (2012) 1044–1053. [54] W.-J. Du, Q. Yin, L. Cheng, Experiments on novel heat recovery systems on rotary
[33] Y. Ai, X. Liu, Y. Huang, L. Yu, (2020) Numerical analysis of the influence of molten kilns, Appl. Therm. Eng. 139 (2018) 535–541.
pool instability on the weld formation during the high speed fiber laser welding, [55] V. Karamarković, M. Marašević, R. Karamarković, M. Karamarković, Recuperator
Int. J. Heat and Mass Transfer 160 (2020), 120103. for waste heat recovery from rotary kilns, Appl. Therm. Eng. 54 (2013) 470–480.
[34] AEA Technology, (2001) CFX-4.4: Solver, AEA Technology, Harwell Laboratory, [56] H. Kramers, P. Croockewit, The passage of granular solids through inclined rotary
Oxfordshire, UK. kilns, Chem. Eng. Sci. 1 (1952) 259–265.

17

You might also like