• Keine Ergebnisse gefunden

PCC-LES M ODEL F ORMULATION

5.1. GAS-PHASE MODELLING 48

5.1 Gas-phase Modelling

5.1.1 LES Transport Equations for the Gas-phase

The basic transport equation for the gas-phase, introduced in section 3.1 and fil-tered by the procedure presented in section 4.3, including basic modelling and clos-ing strategies, are discussed next.

The Favre filtered continuity equation Eq. (4.6) including source terms to reflect mass transfer between the particle- and the gas-phase can be written as

∂ρ

t+

xi

¡ρuei¢

=X

nn1, (5.1)

where P

nn 2 is the mass transfer term between the dispersed and gas-phase ac-cording to a sub-model n, including contributions from pyrolysis and char conver-sion. Additionally, the transferred mass from the particle- to the gas-phase can include evaporation of moisture from coal. In this work, however, evaporation has been neglected because of the low moisture content of the considered coals. All considered source terms are summarised in Tab. 5.1.

Starting from the Favre filtered momentum equation Eq. (4.10) and applying the Boussinesq approximation, cf. Appendix A.3 for details, gives

∂ρuei

t +

xi

¡ρueiuej¢

=

xi

µEff Ãuj

xi +ui

xj −2 3

uk

xkδi j

!

−p

xi +X

n Λin3, (5.2) where P

nΛin is a momentum source considering gravitation and momentum transferred from or to the particle-phase by relative velocities between parti-cles and gas-phase. The filtered effective stress tensor τEff is approximated as µEff

³

uei

∂xi +∂xuejj23∂xuekkδi j

´ with the effective viscosity µEff =µ+µSGS. The filtered molecular viscosityµ is approximated using the filtered cell temperature based on the following equivalent of Sutherland’s law as

µ= Asp T

³1.0+Ts/T´, (5.3)

1For the OpenFOAM implementation of this Eqn see listing Lst. C.4

2The mass transfer term corresponds toP

nn=Part from Eq. (4.4) and Eq. (4.6), here written as summation term to indicate that different sub-model contributions are considered individually

3cf. listing Lst. B.1

where As and Ts are mixture dependent coefficients. The SGS viscosity µSGS is computed by the LES turbulence model, which is discussed in detail in section 5.1.2.

The Favre filtered transport equation for the mass fraction of specieskis derived from Eq. (3.7), using the gradient assumption to model the unresolved scalar flux

ρu

:

iSGSYkSGS= −µSGS

ScSGS

Yfk

xi, (5.4)

and reads

∂ρYek

t +

xi

¡ρueiYek¢

=

xi õ

µ

Sc+ µSGS ScSGS

Yek

xi

! +X

n Ψkn 4, (5.5) whereP

nΨkn is the species source term5, Sc, and ScSGS are the molecular and tur-bulent Schmidt numbers, respectively. The composition of the gas-phase is affected by homogeneous reactions and by species transfer between the coal particles and the gas-phase. The latter originates from pyrolysis reactions and char conversion.

To ensure that the sum of species mass fractions equals unity, Eq. (5.5) is solved for k−1 species only. The mass fraction of the inert N2species is then calculated from the remainder of the sum of reacting species mass fractions

YN2=1−X

k−1Yk. (5.6)

Starting from the simplified transport equation for sensible enthalpy Eq. (3.6), using the gradient assumption to model the unresolved enthalpy flux terms in anal-ogy to the unresolved filtered species flux term, yields

∂ρheS

t +

xi

¡ρueiheS¢

=

xi µ

cpαEffTe

xi

¶ +X

n Θn, (5.7)

where the cumulative enthalpy source term is expressed as P

nΘn. The filtered effective heat diffusion coefficientαEffis calculated from the molecular and the SGS viscosity using the following relation

cpαEff=cp¡

α+αSGS¢

=cp µ µ

Pr+µSGS Prt

λEff6, (5.8)

4cf. listing Lst. C.7

5Here, again considering the contributions of then-th sub-model, thusP

nΨkn=Ψk,Hom+Ψk,Part

in Eq. (3.7).

6Here, the OpenFOAM convention that the density is absorbed into the heat diffusion coefficient is used. Thus, Eq. (5.8) is written in the formcpα=λinstead ofcpα=λ/ρ.

5.1. GAS-PHASE MODELLING 50 where Pr and PrSGS denote the laminar and turbulent Prandtl numbers. For better stability and to reduce computational costs the diffusion term in Eq. (5.7) is approx-imated by αEff∂xheS

i

7. Additionally, the enthalpy diffusion induced by species gradi-ents ∂x

i

³ ρP

khk,SYkVik´

has been neglected in Eq. (5.7). Thus the filtered sensible enthalpy transport equation can be written as

∂ρheS

t +

xi

¡ρueiheS¢

=

xi Ã

αEffheS

xi

! +X

n Θkn. (5.9)

For better readability source terms in the transport equations have been pre-sented by a summation term. A detailed list of individual source terms is listed in Tab. 5.1.

Table 5.1: Individual transport equation source terms.

Term Comment Section

Mass Source TermsP

nn,Eq. (5.1)

CC Particle char conversion 5.3.5

Dev Particle devolatilisation 6.2

Momentum Eqn. Source TermsP

nΛin,Eq. (5.2)

gi Gravitational force

i,Dr Particle drag forces 5.3.2

Λi,Disp Particle dispersion forces 5.3.2

Species Source TermsP

nΨsn,Eq. (5.5)

Ψk,Hom Homogeneous gas-phase reactions 5.1.3

Ψk,CC Particle char conversion (e.g. CO, CO2) 5.3.5 Ψk,Dev Particle devolatilisation (e.g. CH4, H2, . . . ) 6.2

Enthalpy Source TermsP

nΘkn,Eq. (5.9)

ΘHom Homogeneous gas-phase reactions 5.1.3 ΘCond,G Conductive heat transfer from coal particles 5.3.3 ΘRad Combined particle and gas-phase radiation 5.3.4 ΘCC Enthalpy of particle char conversion reactions 5.3.5 ΘPy Enthalpy of particle pyrolysis reactions 6.2

7Using the following relation

xi µ

cpα∂T

xi

=

xi µ

αcpT

xi αTcp

xi

,

and replacingcpT=hS, it becomes clear that the termTcp/xihas been omitted in Eq. (5.7).

5.1.2 Turbulence Model

In the process of creating the results presented in the chapters 7 and 8, several tur-bulence models have been employed. For turbulent flows which are not bounded by walls, cf. Chap. 7 or non-resolved wall modelled flows, cf. Chap. 8, the OpenFOAM Smagorinsky-Lilly type model was applied. Theσ-model, Nicoud et al. (2011), was used for simulations of pipe flows with the purpose of creating LES inflow data, having resolved near wall flows.

Smagorinsky-Lilly Type Turbulence Models

This section discusses the procedure which is used to compute the turbulent viscosi-ties and its relation to the model proposed by Lilly (1962) and Smagorinsky (1963), which is commonly known as Smagorinsky-Lilly or Smagorinsky model. Turbulent viscositiesµTurb, e.g.µSGS, needed for the closure of Eq. (5.2), Eq. (5.5), and Eq. (5.9) are often defined based on dimensional grounds. For example, Tennekes and Lum-ley (1972) give the following relation based on a mixing lengthlTurb

µTurbρuTurblTurb, (5.10) whereuTurb is a turbulent velocity scale.

The relation Eq. (5.11), formulated by Prandtl and Wieghardt (1946) µTurbρlTurbq

kTurb, (5.11)

wherekTurb8is the turbulent kinetic energy, was a precursor for modern one equa-tion models of turbulence, Speziale (1991). This approach, of using the turbulent kinetic energy, is more flexible since a transport equation for the turbulent kinetic energy can be used to solve forkTurb. This can increase the accuracy while increas-ing the computational costs. In the LES context this can be formulated as

µSGS=ckρqkTurb, (5.12) where lTurb is replaced by the product of a model constant ck and the grid spacing

∆. Another way of representing µTurb uses the turbulent kinetic energy kTurb and its dissipation rate² as discussed by Menon et al. (1996) or Yoshizawa (1987).

8The turbulent kinetic energykTurbis a mass specific quantity and has the units m2s−2.

5.1. GAS-PHASE MODELLING 52 To formulate an equation for kTurb a local equilibrium of the influx of turbu-lent kinetic energy from the resolved scales and its dissipation ² is assumed. The equilibrium can be written as

τi j,SGS:Sei j=²(∆), (5.13)

where τi j,SGS models the subgrid scale viscous stress tensor, cf. Eq. (4.9), and : denotes the Frobenius inner product, cf. Eq. (A.1). The grid spacing dependent cut off dissipation rate²(k) can be written as

²(k)=c²ρk3/2Turb−1, (5.14) with the model constant c², Fureby (1996). The subgrid scale viscous stress tensor τi j,SGS is modelled by the Boussinesq approximation. After inserting Eq. (A.9) and Eq. (5.14) into Eq. (5.13) and replacingµTurb with Eq. (5.12) the algebraic equation for kTurb can be written as

−c²ρk3/2Turb−1=ρ2

3kTurbδi j:Sei j+2ρckk1/2Turb∆Sei j,Dev:Sei j, (5.15) or reformulated

c²

ρkTurb+ 2

3ρk1/2Turbδi j:Sei j−2ρck∆Si j,Dev:Sei j=09, (5.16) where Eq. (5.16) is a quadratic equation forp

kTurb andSi j,Devthe deviatoric strain rate tensor, cf. Eq. (A.9).

In the constant density case the transport equations can be divided by ρ and µTurb/ρsubstituted byνTurb, Eq. (5.15) is then be simplified to

kTurb=2ck c²

2S2i j, (5.17)

sinceδi j:Sei j=0, as a consequence of the continuity equation10.

By inserting Eq. (5.17) into Eq. (5.12) the commonly used Smagorinsky constant CSmagcan be expressed as

CSmag= Ãc3k

c²

!1/4

. (5.18)

9cf. listing Lst. B.9

10Additonally, the relation Sei j,Dev:Sei j=Se2i j holds, since Sei j,Dev=Sei j for incompressible flows.

This follows from the definition given by Eq. (A.7)

This demonstrates that the standard Smagorinsky-Lilly model, i.e.

µSGS=ρ¡

CSmag¢2¯¯ eSi j¯

¯, (5.19)

can be recovered in the constant density case from the generalised variable den-sity formulation based on the local equilibrium assumption for the kinetic energy.

Hence, the model based on Eq. (5.12) in combination with the quadratic equation for pkTurb, Eq. (5.16), is referred to as generalised compressible Smagorinsky model.

σ-Model

An alternative LES turbulence model, used within this work, is the σ turbulence model, proposed by Nicoud et al. (2011). Theσcalculates the turbulent SGS contri-butions to the viscosity as

νSGS

Cσ¢2σ3

¡σ1σ2¢ ¡

σ2σ3¢

σ21 , (5.20)

whereσi are the square roots of the Eigenvalues of the matrixGi j=∂u∂xki ∂u∂xkj.

5.1.3 Turbulence Chemistry Interaction Models

Since the instantaneous concentrations and the temperature needed for Eq. (3.14) are generally not known in a turbulent flow and applying mean or filtered values instead would lead to inaccurate mean reaction rates, discussed in appendix A.5, an approach to solve this fundamental problem of turbulence chemistry interaction (TCI) is needed.

For the closure of the LES species transport equation, the filtered reaction rate Ψk,Hom is needed. In the present work the eddy break-up (EBU) model of Hu et al.

(2006) has been implemented. The basic assumption of the EBU model are infinitely fast chemical reactions in comparison to the turbulent mixing, i.e. τChem<<τMix. Hence, turbulent mixing is the factor that controls the reaction rate. The filtered fuel consumption rateΨFuis defined as

ΨFu=c1ρ¯

¯ eSi j¯

¯min Ã

YeFu,YeOx

s ,c2 YePr 1+s

!

, (5.21)

whereYeOx,YeFuandYePrare the filtered mass fractions of the oxidiser, fuel, and prod-ucts. Furthermore, s=¡

νOxMWOxFuMWFu¢

St is the mass stoichiometric factor.

5.2. SOLID PHASE MODELLING 54 The default values of the model coefficients c1andc2are 4.0 and 0.5. After evaluat-ing the fuel consumption rate, the individual species source terms can be calculated by Ψk,Hom= ηηk

FuωFu, where ηk =νStMWk is the mass stoichiometric coefficient of the species k and ηFu the mass stoichiometric coefficient of the corresponding fuel species.

OnceΨk,Hom is known, Θn,Hom, needed for the closure of the sensible enthalpy transport equation, can be calculated.

Θn,Hom=HVFuΨFu,Hom, (5.22)

where HVFu is the heating value of the fuel for the corresponding reactionn.

Besides its low computational costs and comparably simple implementation, the EBU model has the advantage that no Arrhenius rate coefficients are needed since the reaction rates are only dependent on the turbulent time scales. This makes the EBU model suitable whenever the exact nature of the reacting species is unknown, e.g. whenever tars or species mixtures are approximated by postulate substances.

5.2 Solid Phase Modelling

The particle position xi,Partchanges according to d

dtxi,Part=ui,Part. (5.23)

Hence, the position of each particle is determined by integrating Eq. (5.23) for each particle individually. The unknown particle velocityui,Partcan be found using New-ton’s second law

mPart d

dtui,Part=Fi,Tot, (5.24)

where Fi,Tot is the sum of various forces acting on the particle and neglecting any momentum changes due to changes in the particle mass. Particles are assumed to be spherical and non-rotating. Thus, in the present study only drag, gravitational, and dispersive forces arising from the modelled turbulent fluctuations are consid-ered,

Fi,Tot=Fi,Dr+Fi,G+Fi,Disp. (5.25)

The dispersive force FDisp in Eq. (5.25) is a force introduced by the filtering proce-dure, reflecting the dispersive nature of the subgrid scale gas-phase velocity fluctu-ations on the particle motion.

An updated particle velocity can be obtained by integration over Eq. (5.24). The implementation of this integration step, however, involves further transformations of the differential equation. The details are discussed in appendix C.3.1.

The particle temperature is calculated by balancing the energy absorbed from the surrounding gas-phase via convection, conduction, radiation, and additionally from processes such as evaporation, pyrolysis, or surface reactions. The resulting equation can be written as

d dt

¡mPartcp,PartTPart¢

Cond,PartRadReac, (5.26) where ΘCond,Part is the conductive heat transfer term between particle- and gas-phase, ΘRad the heat transferred by radiation, and ΘReac the heat released due to chemical processes. The conductive heat transfer termΘCond,Partis given by

ΘCond,Part=NuπdPartλEffG

³TG−TPart´

, (5.27)

where Nu is the Nusselt number andλEff the effective thermal conductivity of the gas-phase.

Equivalently to the particle motion equation the particle temperature equation is subject to further transformation discussed in appendix C.3.2.

The final, fundamental particle property that needs to be known is the change in particle mass, which can be expressed as

dmPart

dt =ΩEv+ΩDev+ΩCC, (5.28)

whereΩEv,ΩDev, andΩCC are the mass transfer terms to account for evaporation, devolatilisation and surface reactions respectively between particle- and gas-phase.

The total mass of a particle is considered as the sum of the individual masses of moisture, fixed matter, and volatile content

mPart=mMoi+mVol+mFM. (5.29) The fixed matter mass component mFM consists of the individual masses of ash, char, and coal in the intermediate state.

mFM=mAsh+mChar+mInt. (5.30)

5.3. INTERPHASE TRANSFER MODELLING 56 Here, intermediate state refers to the particle mass which is convertible to char or volatile matter by pyrolysis reactions. Initially, a particle consists only of mIntand mAsh since mVol(t=0)=mChar(t=0)=0. mVol and mChar are then formed by the thermolysis model. The mass of volatiles mVol specifies the mass of species which are potentially released by devolatilisation, but are not emitted to the gas-phase yet. The default implementation in OpenFOAM requires an a priori specification of mVol. Here the author introduced the thermolysis model which allows to form volatile species from the intermediate state. The change of mIntis given by

dmInt

dt = −dmChar dt

¯

¯

¯thermolysis−dmVol dt

¯

¯

¯thermolysis. (5.31) Hence char and volatiles are formed by consuming mInt. The actual form of Eq. (5.31) depends on the selected submodel, discussed in detail in section 5.3.5 and the implementation in appendix C.5.1.

5.3 Interphase Transfer Modelling

This section focuses on the presentation of the transfer terms between particle- and gas-phase. For each source in the gas-phase equations given in Tab. 5.1 a sink in a particle equation exists, i.e. ΦG = −ΦPart and vice versa. In the context of the discretised transport equations a general source term for the cell i can, thus, be written as

Φi,G= − 1 Vi

X

n Φn,Part (5.32)

where nis the number of the particles inside the cell i.

Whenever possible the transfer models are generalised so that instead of the source or sink term a model variable is calculated. For example, instead of comput-ing theΘCond,Part the model for the conductive heat transfer computes the Nusselt number, Nu. In a second step, outside the scope of the transfer model, the particle and gas-phase source and sink terms are calculated. This detail shall be stated here to allow for a better understanding of the implementation of the individual sub-models. Table 5.2 gives an overview of the most relevant sub-model implemen-tations. Here, the aforementioned variable which is solved rather than the source or sink term is given in the “particle model term” column. Once the particle model has been evaluated, in some cases several, dependent, gas-phase source or sink terms can be evaluated. These are given in the “gas-phase source term” column in the Tab. 5.2.

Table 5.2: Coal particle sub-model terms Sub-model Particle Model

Term Gas-phase Source

Term Implementation

Momentum Transfer Models, section 5.3.2

Sphere drag cDr Λi,DrEq. (5.2) •Schiller and Naumann (1933) Particle

dispersion Fi,Disp Λi,DispEq. (5.2) •No model

•Bini and Jones (2008) Heat Transfer Models, section 5.3.3 and 5.3.4

Conductive

heat transfer Nu ΘCond,GEq. (5.7) •Ranz and Marshall (1952) Radiation EPart,κPartRad,Part ΘRadEq. (5.7) •P1with fixed constants

Particle Reaction Models, section 5.3.5 and section 6.2 Char

conversion dtdmChar

CCEq. (5.1), Ψk,CCEq. (5.5), ΘCCEq. (5.7)

•Baum and Street (1971)

Pyrolysis dtdmVol

DevEq. (5.1), Ψk,DevEq. (5.5), ΘPyEq. (5.7), ΘReacEq. (5.7)

•Badzioch and Hawksley (1970)

•Kobayashi et al. (1977)

•Richards and Fletcher (2016)

5.3.1 The Parcel Approach

To reduce the computational costs individual particles with identical diameter are combined to parcels. As a consequence all particles of a parcel are identical and thus all properties, e.g. velocity, position, temperature, and diameter are the same.

Source terms transferred to the gas-phase, e.g. mass and energy, are calculated as the product of the number of particles per parcel and the individual source term, i.e. ΩParc=nPartPart. This approach is easy to implement since all submodels are formulated on a per particle basis and only the source terms need to be adapted by the total number of particles per parcel. A drawback of this method is that for a large number of particles per parcel on fine computational grids, the source terms are concentrated in a limited number of cells. This can reduce the accuracy of the simulation and potentially cause instabilities.

5.3.2 Particle Momentum Transfer Models

Particle Drag Model

The relevant particle sub-model term to model momentum transfer between parti-cle and gas-phase is the partiparti-cle drag coefficient cDr. In this work, the drag

coeffi-5.3. INTERPHASE TRANSFER MODELLING 58 cient of a coal particle is modelled with a modified version of the Reynolds number dependent correlation of Schiller and Naumann (1933)11

cDr=





0.424 if RePart>1000

Re24Part

µ

1+16RePart23

else , (5.33)

where RePartis the particle Reynolds number defined as RePart=uReldPart

νG , (5.34)

where uRel is the relative velocity between particle and gas-phase and dPart the particle diameter. Equation (5.33) considers two distinct flow regimes. At high Reynolds numbers, based on the relative particle velocity, a constant particle drag coefficient value of 0.424 is considered. In the low Reynolds number regime a mod-ified version of the Stokes relation, cDr=24/RePart, with a multiplicative correction factor is considered.

After calculating the particle drag coefficientcDr the particle drag force for the n-th particleFin,Drand therefore the gas-phase source term can be evaluated as

Λi,Dr= −1 V

X

n Fin,Dr, (5.35)

with

Fin,Dr=1

2ρGAn,Cr,Partu2i,RelcDr, (5.36)

where An,Cr,Partis the cross sectional area of a sphere.

Particle Dispersion Model

To reflect the dispersive nature of the subgrid scale fluctuations the LES particle dispersion model of Bini and Jones (2008) has been implemented in this work. The model of Bini and Jones (2008) calculates a dispersion velocity as

dui,Disp= s

C0kSGS

τt dWi, (5.37)

whereτt is the interaction timescale between particles and the unresolved vortices, kSGSthe kinetic energy of the subgrid fluctuations,Coa coefficient, andWa Wiener

11This is the default drag coefficient for spherical particles of OpenFOAM.

term. For the implementation of the dispersion model within the framework of OpenFOAM, a dispersive force acting onto a Lagrangian particle is calculated by applying Newton’s second law to Eq. (5.37)

Fi,Disp=mPartdui,Disp

dt . (5.38)

SubstitutingdvDispin Eq. (5.38) with Eq. (5.37) yields

Fi,Disp=mPart r

C0kSGSτt dWi

dt . (5.39)

For the interaction timescale τt multiple definitions are given by Bini and Jones (2008). In this work, however, only the following relation is used

τt= τPart

¡∆/pkSGS¢2α−1, (5.40)

where∆is the cell width,τPart the particle timescale, and the coefficientαis set to 0.8.

Analogously to Eq. (5.35) the momentum transfer term for the particle disper-sion force can then be written as

Λi,Disp= −1 V

X

n Fin,Disp. (5.41)

5.3.3 Particle Heat Transfer Model

The Ranz and Marshall Model

Based on the work of Frössling et al. (1938), the model of Ranz and Marshall (1952) estimates the Nusselt number, which is needed for closure of Eq. (5.27), as a func-tion of the Reynolds and Prandtl number as

Nu=2+0.552Re1/2PartPr1/3. (5.42) Even though developed to model the heat transfer of evaporating droplets, the model of Ranz and Marshall (1952) is also widely employed to compute the con-vective heat transfer of coal particles.

After calculating the particle Nusselt number Nu the conductive heat transfer terms for the individual particles Eq. (5.27) can be solved for. In consequence the

5.3. INTERPHASE TRANSFER MODELLING 60 heat transfer for term for the gas-phase can be calculated as

ΘCond,G= −1 V

X

n ΘCond,Part (5.43)

5.3.4 Radiation Modelling

A number of methods to solve the RTE Eq. (3.12) have been developed. These in-clude: discrete ordinates methods (DOM), implicit Monte Carlo (IMC), and Flux Methods to name only a small selection. Due to the enormous amount of different methods and their complexity these solution methods cannot be presented in detail.

The interested reader is referred to the review of Viskanta and Mengüç (1987) or the book of Modest and Haworth (2016) for further details. The approach of choice to solve the RTE in the present report is the spherical harmonics (PN) method of order n=1.

The P1Method allows for the transformation of the RTE, Eq. (3.12), which is an integro-differential equation with five independent variables, into a set of regular ODEs depending only on spatial coordinates. The full procedure of deriving the transport equation for the incident radiation G, i.e. G =R

IdΩ, from the RTE, which is based on Legendre polynomials, is rather lengthy. Therefore, the interested reader is referred to Modest (2003) for a detailed discussion.

The resulting P1transport equation for the incident radiation is given as

xi µ1

β

G

xi

= −3κ¡

4πIBB−G¢

, (5.44)

where βis the extinction coefficient. Furthermore, from Eq. (5.44) the source term for the sensible enthalpy transport equation is given as

ΘRad=

xi µ1

β

G

xi

. (5.45)

To account for particle contributions in Eq. (5.44) total absorption and extinction coefficients are used. Further details on this and the corresponding implementation are given in appendix B.7.