Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 23

The University of Birmingham

School of Chemical Engineering

Year 2 Bridging Coursework

A Study of the Biomass Gasification Process

Module Codes: Mass, Heat and Momentum Transport (04 17125)


Process Integration and Unit Operations (04 17126)
Process Systems and Principles of Process Control
(04 17123/04 20490)
Chemical Engineering Thermodynamics (04 22992)

Irfan Zhariff Bin Hazali


ID 2095684

UG Chemical Engineering (Affiliated)

June 2020

ABSTRACT:
Globally, there has been increasing interest in biomass as a source of alternate
energy. One of the routes to generate energy from biomass is biomass gasification.
Carbonaceous feedstock is converted into syngas, an intermediary fuel that can be
used to produce other gases such methanol, ethanol, and hydrogen. This work
analyses the biomass gasification process from various angles: heat transfer, mass
transfer, thermodynamics, process control and process integration, and touches
upon applications of liquid-liquid extraction and membrane separation. It illustrates
that knowledge gained from the first two years of the programme allows students to
begin a comprehensive analysis of a practical Chemical Engineering problem.
Please note that as an exchange student, I did not take Sustainable Process
Engineering (04 31633) and Reactors, Catalysis and Thermodynamics (04 32470). I
have taken Chemical Engineering Thermodynamics (04 22992) from Year 3, and so
those learning outcomes will be targeted as well.

Table 1: List of targeted learning outcomes


Learning Outcomes
MHM1 Use fundamental transport phenomena understanding to generate and
simplify relevant engineering problems.
MHM2 Use fundamental understanding of transport phenomena mathematics
to solve relevant engineering problems.
PIU1 Apply problem table and energy cascade to determine the minimum hot
and cold energy requirements and the pinch point of a heat exchange
system;
PIU2 Design a heat exchanger network for maximum energy recovery or
minimum number of exchangers; comment upon the appropriate use of
process integration in designing new chemical and process plant and
revamping existing plant;
PIU3 Demonstrate and apply the fundamentals of the major unit operations
in Chemical Engineering namely distillation, extraction and
crystallization; in terms of the essential requirements for the unit
design, detailed calculations to find the number of stages and/or the
size of unit needed to perform a certain function;
PIU4 Explain the principles of supercritical fluid technology, in terms of the
supercritical fluids, properties, and the main processes that are based
on these properties;
PIU5 Describe and review the principles and applications of membrane
treatment systems and size membranes for single or feed and bleed
stage systems;
PIU6 Appropriately select a unit operation for a particular process need.
PSPPC1 Appreciate the importance of process control and the role of the
process control engineer particularly in safety critical systems;
PSPPC3 Describe primary sensing elements used for commonly measured and
manipulated process variables select the appropriate final control
element for a given process application and how these are used within
control loops on process plant;
CET1 Describe and use relevant equations of state. (Assessed in coursework
as well)
CET2 Demonstrate the importance of Chemical Potential and Fugacity.
CET3 Relate properties to important thermodynamic cycles and use these in
practical examples. (Assessed in coursework as well)
CET4 Discriminate different forms of equilibria and if phase separation could
occur.

1
Contents
1. Introduction............................................................................................................2
2. Drying.....................................................................................................................3
3. Gasification............................................................................................................4
4. Energy Balance......................................................................................................6
5. Mass Transfer Coefficient......................................................................................9
6. Heat Transfer.........................................................................................................9
7. Equilibrium Model................................................................................................12
8. Process Integration..............................................................................................14
9. Process Control....................................................................................................16
10. Liquid-Liquid Extraction....................................................................................17
11. Membrane Separation......................................................................................17
12. Conclusion........................................................................................................18
13. Appendix...........................................................................................................19
14. References.......................................................................................................21

1. Introduction
The rapid industrialisation of the 20th century came with a global change in lifestyle:
people became increasingly dependent on cheap, accessible energy – the vast
majority of which were derived from non-renewable fossil fuels such as petroleum,
coal, and natural gas. This dependency directly impacted carbon dioxide emissions
to the atmosphere, increasing effects of climate change by encouraging global
warming.

As these resources deplete, alternative sources of energy are continually being


investigated. This includes deriving energy from waste. Biomass is defined as
organic matter derived from plant or animal waste, biomass is mainly composed of
carbon, hydrogen and oxygen, and already exists unused in large quantities as
forestry waste and agricultural by-products.

Converting biomass to energy can be done through biochemical or thermochemical


means(Chandra and Boravelli, 2016). The former produces fuels through anaerobic
digestion of biomass; the latter includes methods such as combustion, pyrolysis and
gasification, each with different final products.

Combustion is the most direct method to generate heat, though the overall efficiency
is low. Pyrolysis results in solid, liquid, or gaseous products. Gasification has the
additional capability of converting biomass into intermediary fuels such as
syngas(Kumar et al., 2009). With minor equipment modifications, these may be able

2
to utilise existing natural gas infrastructure or be utilised themselves in fuel cells to
raise efficiency of power production(Ptasinski, 2015).

The main stages in the process of biomass gasification is summarised in Figure 1-1.

Figure 1-1: Biomass gasification process (Kumar et al., 2009)

Any fluid beyond the critical point (T and P ), when gaseous and liquid phases
C C

coexist as a single phase, is described as “supercritical”. Fluids at this point have


interesting properties: their density is very sensitive to pressure changes, so to
separate and isolate materials from solution, one needs only to vary the pressure;
this causes the target material to precipitate due to differences in solubility. The gas
phase density becomes similar to the liquid phase density; this large increase from
normal densities in the gas phase means that the fluid has much greater solvating
power. Diffusion and mass transfer limitations are eliminated as diffusion coefficients
are very large (close to gas diffusivity). 

Supercritical water gasification (SCWG) is an emerging technology that permits high


solid conversion (and so low levels of tar and char). Water, being low-cost and non-
toxic, becomes supercritical past 374C and 22.1MPa (Antal et al., 2000). Beyond this
point, hydrogen bonds between molecules are weakened, contributing to H 2

production. Thus, supercritical water has a dual role as a medium and reactant in the
gasification of biomass; H and OH ions are generated at greater densities,
+ -

contributing to an environment for hydrolysis and pyrolysis reactions. Additionally, at


supercritical conditions, cellulose and hemicellulose (a polysaccharide with a simpler
structure than cellulose) are dissolved by water, since it acts as a strong oxidising
agent(D. et al., 2015). Instead of yielding conventional syngas, SCWG produces H 2

and CO , which is easily separated from water by cooling at room temperature.


2

2. Drying
After the biomass has been reduced in size (usually through knife mills, hammer
mills, or tub grinders), it needs to be dried. Decreasing the water content of the
biomass increases the overall efficiency of the process, decreasing electricity cost.
The amount of heat required depends on the moisture content of the feedstock, but
typically begin from 10 to 35 wt% and is reduced to 5 wt%. Rotary cascade,
perforated bin and band conveyor dryers have been used to dry biomass (Cummer
and Brown, 2002).

3
The following case study is from (Shrestha, 2014).

Table 2-1: Parameters for drying case study

Wet biomass flowrate, mwet (kg/hr) 27.85

Moisture content, Mc (wt%) 10.12

Carbon in feedstock, Cin (kg/hr) 25.03

Carbon content, c (%) 51.13

Molar fraction of CO, YCO 0.256

Molar fraction of CO2, YCO2 0.098

Molar fraction of CH4, YCh4 0.015

Molar density of ideal gas at standard 44.615


temperature and pressure, ρmol (mol/m3)

Molecular weight of carbon, RC (kg/mol) 12

Actual flowrate of syngas, F (m3/hr) 55

To calculate the flowrate of dry biomass, the amount of moisture m c is subtracted


from mwet. Cin is calculated by multiplying mdry with c.

10.12
m dry =m wet −m c =27.85− ( 100 )
( 27.85 )=25.03 kg /hr

C ¿=25.03∗ ( 51.13
100 )
=12.79 kg /hr

The total carbon leaving with the syngas, C out, is calculated by equation X.

C out =( Y CO +Y CO 2+Y CH 4 ) ∙ ρmol ∙ R C ∙ F

12
C out =( 0.256+ 0.098+0.015 ) ∙ 44.615∙ ∙55=10.87 kg /hr
1000

3. Gasification
Gasifiers can be categorised by three main types of reactor design: fixed-bed,
fluidised bed, or entrained flow. This is illustrated in Figure 3-1.

4
Figure 3-1: Types of gasifiers (Ptasinki, 2016)
In fixed-bed gasifiers, the solid biomass particles move down the reactor vessel, and
the gasifying agent either moves co-currently (downdraft) or counter-currently
(updraft). They operate at 1000-1200°C (downdraft) or 800-1000°C (updraft), and
the residence time of the biomass is 15-30 minutes, ensuring high carbon
conversion. Downdraft gasifiers produces syngas with low tar content (0.015-0.5
g/Nm3), whilst for updraft gasifiers the tar content is high (30-150 g/Nm 3).
Fluidised-bed gasifiers suspend the biomass particles by pumping the gasifying
agent at velocity. Very good mixing occurs, resulting in uniform temperature
distribution and very short residence times (5-50 s). Unlike fixed-bed gasifiers,
separate reaction zones cannot be distinguished (Figure 3-1c and 3-1d). The tar
content is 0.1-30 g/Nm3 and they operate at 800-900°C(Molino et al., 2016).
Within gasification, there are four main stages: oxidation, drying, pyrolysis, and
reduction. The reactions for each are described below (Ptasinki et al., 2009):

Oxidation: C+ O2 → CO2 ∆ H=−394 kJ /mol (R1)

1 (R2)
C+ O 2 →CO ∆ H=−111 kJ /mol
2

1 (R3)
H + H 2 → H 2 O ∆ H=−242 kJ /mol
2

Drying: Moist Fuel → Fuel+ H 2 O(g) (R4)

Pyrolysis: Fuel ⇌ H 2+CO +CO 2+CH 4 + H 2 O ( g )+ Tar+Char (R5)

Reduction: C+ CO2 ⇌ 2 CO ∆ H =172kJ /mol (R6)

C+ H 2 O ⇌CO+ H 2 ∆ H =131 kJ /mol (R7)

CO+ H 2 O ⇌ CO 2+ H 2 ∆ H =−41 kJ /mol (R8)

5
C+ 2 H 2 ⇌CH 4 ∆ H=−75 kJ /mol (R9)

The most relevant are reactions (2), (6), and (8), which yield the most carbon
monoxide and hydrogen, the main constituents of syngas. The global chemical
reaction may be written as:
C x H y O z+ a H 2 O+ Heat ⟶ bCO+ dC O 2+ e H 2 +fC H 4 + g H 2 O Eq .(3−1)

Gasifier design depends on some of these factors(Kumar, Gupta and Viswanadham,


2018):

1. Fuel Characteristics - Energy content, moisture content, size, shape and


density of feedstock all play a part in the design of the gasifier. If the moisture
content is below 20%, downdraft designs are more suitable; feedstocks with
higher fuel densities require smaller reactor size. 
2. Operating Temperature - Efficiency increases at higher temperatures with
energy losses increasing as well. Insulation is required to reduce energy
losses from the reactor chamber.
3. Residence Time - Longer residence times decreases formation of tar,
increases carbon conversion, and improves gas yield.
4. Superficial Velocity - The ratio of syngas production rate and cross-sectional
area of the gasifier. Depends on feedstock packing factor, which creates
resistance to air flow. Low superficial velocity corresponds to slow pyrolysis
process and high amounts of tar and char.
5. Reactor Cross-sectional Area - Depends on fuel consumption and gasification
rate.

6. Design Considerations - Height of reactor affects operational time and gas


volume produced from reactor column. Greater height increases resistance to
air flow, requiring a more powerful draught system.

4. Energy Balance
The following is based on (Arizaleta, 2018). The overall energy balance can be
written as below:
Accumulation=Heat added−work done +energy ∈−energy out

In terms of defined variables this can be written as:


dE
=Q̇−Ẇ S + Ė mass∈¿− Ė Eq .(4−1)¿
dt mass out

Where Ėmass∈¿∧Ė mass out ¿ can be considered zero in a closed system and so the equation
reduces to:
dE
=Q̇−Ẇ S
dt

6
In closed systems, the kinetic and potential energy is neglected, causing the internal
energy, U, to be the main energy contribution. This means equation X can be written
as:
dU
=Q̇−W˙ S Eq .(4−2)
dt

Internal energy can be expressed in terms of enthalpy.

U =∑ M i ui=¿ ∑ M i ( hi −P v i ) Eq .(4−3)¿
i i

Where P = system pressure, 𝑢𝑖 = specific internal energy of component , ℎ𝑖 =


specific enthalpy, 𝑣𝑖 = specific volume and 𝑀𝑖 = mass. Equation X can continue to be
manipulated to Eq X, and then substituted into Eq. (X)

U =∑ M i hi −P ∑ M i v i=∑ M i hi−P V Eq .(4−4)


i i i


( ∑ M h −PV ) =Q̇−Ẇ
i
i i

Eq. (4−5)
S
∂t


( ∑ M h −PV ) = ∂ (∑ M h ) − ∂ ( PV ) Eq .(4−6)
i
i i
i
i i

∂t ∂t ∂t
∂ hi ∂ Mi ∂V ∂P
¿∑ Mi + ∑ hi −P −V Eq .(4−7)
i ∂t i ∂t ∂t ∂t

∂V
Using the following correlations, and that P =0 in a closed system, the energy
∂t
balance becomes:
∂h ∂ Mi
∑ M i ∂ t i =∑ M i C p ,i ∂∂tT ∧∑ hi ∂t
=V ∑ Δhk r k
i i i k

∂T ∂P
∑ M i C p , i ∂t +V ∑ Δ h k r k −V ∂t
=Q̇−W˙ S Eq .(4−8)
i k

Where 𝑐𝑝, = specific heat capacity of component 𝑖, ∆ℎ𝑘 = heat of reaction 𝑘 and 𝑟𝑘 =
rate of reaction 𝑘1-3 calculated as in section 5. W S can be considered to be zero since
the system is closed, and generally no movable equipment present. Q is assumed to
be only through conduction which is:
k
Q̇= A ( T S−T ) Eq .(4−9)
tw

Where 𝑘 = thermal conductivity, 𝑡𝑤 = thickness of the stainless-steel wall, 𝑇𝑠 =


temperature of the outer wall, and 𝑇 = temperature inside the reactor. This means
Eq. (X) becomes, once reorganized:

7
∂T ∂P k
∑ M i C p , i ∂t −V = A ( T S−T ) −V ∑ (−Δ hk )r k Eq .(4−10)
∂ t tw
i k

(De Blasi, 1996) Pressure is assumed to follow the ideal gas law.
ng RoT Mg
P= n g=
Vg Wg

Mg = mass of total volatiles, and W g = molecular weight of the total volatiles, so n g =


moles of total volatiles. The reactor volume is the sum of the volume of volatiles V g
(tar and non-condensable gases) and the solid V s.
V =V g +V s

Vs decreases with time, as solid mass is lost during the process.

Vs Ms (Mw+MC)
≈ = Eq .( 4−11)
V s ,i M s ,i Mw,i

Vs,i = initial solid volume, Vs = solid volume at time t, Ms,i = initial solid mass, Ms =
solid mass at time t, Mw = mass of wood, Mc = mass of char. At time t = 0, the reactor
will only contain wood, and so Ms = Mw.
Then, equation X can be rearranged and substituted into Eq. (X).

( M w+ M C )
V g=V −V s=V −V s ,i Eq.(4−12)
M w ,i

Eq. (X) and Eq. (X) can be then substituted into the ideal gas law above.
Mg
R T
Wg o
P= Eq .( 4−13)
( M w+ M C )
V −V s ,i
M w ,i

The pressure derivative with time is then taken, reorganised, and substituted back
into the overall energy balance. This ultimately leads to the final temperature
derivative with time:

1 ∂ Mg Mg V w,i ∂ Mw ∂ Mc

∂T
∂t
=
k
tw
A ( T S −T ) +V ∑ (−Δh k ) r k +V Ro T
k
[
Mg
(
R V
( Wg ∂ t
M +MC
V −V w ,i w
M w ,i
) ( ) (
+

) (
W g M w ,i ∂t

V −V w ,i
+
∂t
Mw+MC
M w ,i
2

)
)
]
Wg o
∑ M i C p , i− M w+ M C
Eq .(4−14 )
i
V −V w , i
M w ,i

8
5. Mass Transfer Coefficient
Chen and Gunkel (1987) developed a model for a downdraft gasification reactor,
following a char particle along a single dimension over any time step Δt.

The superficial gas velocity is calculated using

P
N g=C g v g Eq .(10) and C g= Eq .(5−1)
RTg

Where N g = total mole flux (mol m -2 s-1), C g = total concentration (mol m-3), v g =
superficial velocity (m s-1), R = ideal gas constant, T g = temperature and P = total
pressure. Combining equations (10) and (11) results in

RTg
v g= (N H O + N CO + N CH + N H + N CO + N N )
P 2 2 4 2 2

Where N = mole flux. The concentrations of CO2 and H2O in the bulk are

N CO NH O
C CO , b= Eq .(12)
2
and CH O, b = 2
Eq .(5−2)
2
vg εb 2
v g εb

Where ε = bed void fraction, which is dependent on the ratio of particle to reactor
diameters (Perry et.al, 2018). In this context, the following is proposed

0.2 ( r p −r o )
ε b=0.5−
rp

Where r o = initial particle radius and r o = particle radius at any gasification zone.

The gasification process may be approximated using the mass transfer coefficient
equation below (Wilke et al., 1975):

DA
k =1.17 ℜ0.585 Sc 0.333 Eq .(5−3)
2 ro

2 ro G μ
Where Reynolds number, ℜ= , Schmidt number, Sc= , D A = molecular
μ ρ DA
diffusivity of component A to the mixture, G = superficial mass velocity (sum of
product of molecular weight and mole flux for all components), μ = fluid viscosity,
and ρ = fluid density.

6. Heat Transfer
When heat is applied to a solid particle of biomass, it first reaches the particle
surface by radiation and forced/natural convection(Koufopanos et al., 1991). For
simplicity, we assume that heat flows in one dimension inside the solid by conduction
only (Fourier’s Law, Eq.1). Eq. 2 represents the energy balance of the system.

9
q r=−k ( ∂∂Tr ) Eq .(6−1)
d (ρ Cp T ) −∂ ρ
=−( ∇ ∙q ) +(− Δ H)( )Eq .(6−2)
dt ∂t

Where qr = heat transmission rate (W/m3), k = coefficient of thermal conductivity


(W/m.K), T = temperature (K), r = radial distance (m), ρ = density (kg/m3), Cp =
specific heat capacity (J/kg.K), q = heat transmission rate (W/m3), Δ H = heat of
reaction (J/kg). Eq. (1) and (2) can be combined to form Eq. (3).

1 ∂T
∂(ρ C T ) r( ) [
∂ r (
∂ r )]
p
p

−∂ e
∂t
=p
+ (−∆ H )
∂r
Eq .(6−3) ( ∂t )
Where rp = is the radial distance taking shape factor into account (2 for a sphere).
Assuming ρ , C p , and k to be constant, Eq.(3) can be simplified to Eq. (4).

∂2 T
ρ Cp
∂T
∂t
=k
[( )( ) ( )]
p
r
∂T
∂r
+
∂r
2
+ (−∆ H )
−∂ p
∂t ( )
Eq . ( 6−4 )

Where p = shape factor. Boundary conditions can be taken, assuming a uniform


particle, surrounded by gas and a cylindrical furnace wall.
t=0 , T ( r , 0 )=T O Eq .(6−5 a)

t >0 , r=0 , ( ∂∂Tr )=0 Eq .(6−5 b)


r =R ,−k ( ∂T∂r ) ¿ r= R =h ( t−t f ) +ϵσ (T 4−T 4f ) Eq.(6−5 c )

Where R = particle radius, t = time, tf = time at the end, ϵ = emissivity coefficient, σ =


Stefan-Boltzmann constant, and Tf = final temperature. Introducing the following
groups, Eq. 4 can be written as Eq.6.

k r αt T −T f −ΔH
α= , x= , τ= 2 , θ= ,Q=
ρ Cp R R T o−T f ρ C p ( T o −T f )

H= ( Rk ) ¿
∂θ p ∂θ ∂2 θ −∂ ρ
∂t
=
x ( )( ) ( ) ( )
∂x
+
∂x 2
+Q
∂τ
Eq. ( 6−6 )

τ =0 , θ ( x , 0 )=1 Eq .(6−7 a)

∂θ
τ > 0 , x=0 , =0 Eq . ( 6−7 b )
∂τ

10
∂θ
x=1 , =−Hθ Eq . ( 6−7 c )
∂τ

Where α = thermal diffusivity (m2/s), To = initial temperature, θ = normalized


temperature, x = dimensionless length, and τ = dimensionless time.
(Koufopanos et al., 1991) describes the reaction scheme as below:

Biomass →1 ( Volatile+Gases )1

Biomass →2 ( Char )1

( Volatile+Gases )1 + ( Char )1 →3 ( Volatile+Gases )2 + ( Char )2 Eq .(6−8)

Reactions 1 and 2 are the primary pyrolysis reactions, whilst reaction 3 is the
potential secondary interaction. Power-laws can represent the kinetic scheme.
dB
=−( K 1+ K 2 ) Bn Eq .( 6−9 a)
dt
d C1 n
=K 2 B − K 3 C 1 Eq .(6−9 b)
dt
d C2
=δ K 3 C1 Eq .(6−9 c)
dt

The initial conditions at t=0 are B=1 , C1=0 ,C 2=0 . Reactions 1 and 2 are assumed
to have the same order, and δ represents the volatile and gases deposited on char
sites due to reaction 3.
The rate of pyrolysis can be estimated as in Eq.(10), where the residual mass W is
given by Eq.(11).

∂ρ
∂t
= ( )( ∂∂Wt ) Eq .(6−10)
1
VO

W =B+ C1 +C 2 Eq .(6−11)

Eq.(7) is a boundary-value problem, which can be transformed into a system of


algebraic equations through a finite difference algorithm. The solution of the system
of algebraic equations can be obtained via trial and error, whilst Eqs.(9a-c) can be
approached via analytical methods.

Table 6-1: Value of Parameters Used in Model


Property Value Unit Source
Wood specific heat C p=1112.0+4.85 T c J/kg.K Wenzl (1970)
Wood thermal conductivity C p=1003.2+2.09 T c W/m.K Wenzl (1970)
Char specific heat k =0.13+3 ( 10−4 ) T c J/kg.K Perry et al.
(1984)
Char thermal conductivity k =0.08−( 10−4 ) T c W/m.K Perry et al.
(1984)

11
Solid emissivity ϵ =0.95 Perry et al.
(1984)
Convective heat transfer kg 1
W/m2 Eichhorn et
coefficient
h=
0.322 ( )
l
Pr 3 ℜ0.5
l al. (1960)
3
Initial solid density 650 Kg/m3 (Koufopanos,
1991)

7. Equilibrium Model
Equilibrium models can be used to analyse thermodynamic limits of reactions in a
system; applying this to biomass gasification allows prediction of conditions that lead
to maximum hydrogen yield. The following is adapted from (Sreejith, Arun and
Muraleedharan, 2013), and focuses on gasification of apricot stones with steam as
the gasifying agent.

Assumptions in the model are as follows:

1. Biomass is dry and ash free.


2. Gasification process occurs in thermodynamic equilibrium.
3. No loss of heat to the environment, leading to adiabatic conditions.
4. Higher order hydrocarbons are neglected (so the product gases are only CO,
CO , H , CH and H O).
2 2 4 2

5. Tar and char are not formed from the reaction.


6. Inside the gasifier, temperature and pressure are uniform.

The total Gibbs free energy of a multi-component reacting system is:


N
T
G =∑ (ni μi )Eq .(7−1)
i=1

Considering the system as real gases, partial molar fugacity can be introduced

fi
o
μi=G i + RTln ( ) fo
Eq .(7−2)

The Gibbs free energy of the reacting mixtures reaches a minimum value at the
equilibrium composition, at constant temperature and pressure. The function that
needs to be minimised then is
N N
fi
T

i=1
o
G =∑ ( ni G )+ ∑ ni RTln
i
i=1
( ( )) fo
Eq. (7−3)

G oi is the standard Gibbs free energy, and can be determined by knowing the
standard enthalpy and entropy at the corresponding temperature. Fugacity is the real
gas pressure and so:

12
f i ϕi yi P
= o
fo fi

N N
ϕ i yi P
T

i=1
o
G =∑ ( ni G )+ ∑ ni RTln
i
i=1 ( ( )) f io
Eq .(7−4)

To determine the coefficient of partial molar fugacity, the Redlich-Kwong (R-K)


equation of state can be used

RT ai
P= − Eq .(7−5)
v−bi v ( v +bi )

Where the R-K constants are

0.08664 R T c ,i m 3
a i=
P c, i T 0.5 mol

0.42748 R2 T 2.5c, i
b i= 0.5
Pa
Pc ,i T

And so the fugacity coefficient can be calculated from 

bi a bi ai Z + Bm
ln ϕ i=
bm
( Z−1 )−ln ( Z−Bm )+ m
[ √ ]
b m RT bm
−2
am
ln (Z ) Eq .(7−6)

To calculate properties and coefficients of mixtures (b , a , B ), as well as


m m m

compressibility factor (Z), the following can be used

b m=∑ y i bi
i

a m=∑ y i ai
i

bm P
b m=
RT

Z=∑ y i Z i
i

The global chemical reaction is required for the computation of compressibility


factors of individual gases. Elemental balances of carbon, hydrogen, and oxygen
from the global chemical reaction (Eq. X) act as the constraints for the optimisation
of the non-linear function.

Carbon: x =b+d + f

Hydrogen : y +2 a=2e+4 f +2 g

13
Oxygen: z + a=b+ 2 d+ g

Non−negativity constraint b , d f ,e , g ≥ 0

To methodology to optimise the function is beyond the scope of this piece, but may
be done through an iterative procedure (such as simulated annealing) in software
such as MATLAB.

8. Process Integration
Process integration is defined as a holistic approach to process design, retrofitting
and operation which emphasis the unity of the process (El-Halwagi, 2006), and is
divided into heat integration and mass integration. The former provides an
understanding of heat utilization in the process and uses this to optimise heat
recovery. The latter analyses the flow of mass within the process and uses this to
optimise species generation.  Rather than optimising unit performance separately,
interactions between different units are considered so that resources are employed
most effectively. This maximises site performance by increasing raw material and
energy efficiency, reducing cost and environmental impact.  

Process Integration is typically employed at the beginning of projects, but can be


applied to existing plants as well; retrofitting old plants using these techniques allows
companies to increase profit whilst operating more sustainably (Hallale, 2001).
Recently, PI methodology has been developing through computer-aided simulation
and mathematical optimisation tools, though its foundations are still rooted in Heat
Integration.

(Addenan, 2010) describes a biomass gasification plant with the following process
flowsheet (Figure 8-1), and parameters (Table 8-1):

Figure 8-1: Process flowsheet for biomass gasification plant (Addenan, 2010)

Table 8-1: Parameters for example biomass gasification plant


Stream Description Supply Target ΔH Specific
Temperature Temperature (kW) Heat
(°C) (°C) Capacity

14
(kW/K)
H1 VAPORS and TOP 1026.85 226.85 0.8910 1.11x10-3
C1 MIX and VAPORS 623.54 1026.85 0.6022 1.49x10-2
C2 H20 and STEAM 24.85 249.85 0.1719 7.64x10-4
C3 SORBENT and MIX 24.85 623.54 0.2265 3.78x10-4
H1 is identified as the only hot stream, while C1, C2, C3, are the cold streams. The
initial Δ T min =10 °C . The problem table is set out below.

Table 8-2: Problem table for example biomass gasification plant

Interval Interval Difference in ∑ C p hot −∑ C p cold Enthalpy Surplus


Temperatur Intervals Si (kW/K) (kW) or
e and Si+1 (°C) Deficit
(°C)
-0.00149 -0.0149 Deficit
1031.85 1 10

1021.85
-0.00038 -0.14946 Deficit
628.54 2 393.31
0.00073 0.273541 Surplus
254.85 3 373.69
-0.000032 -0.00106 Deficit
221.85 4 33
-0.001142 -0.21926 Deficit
29.85 5 192

The data from Table 8-2 can be processed into an energy cascade diagram as
below (Table 8-3).
Table 8-3: Energy Cascade Diagram for example biomass gasification process (right
is corrected version)
Interval Enthalpy Total Enthalpy Total
(kW) Enthalpy (kW) Enthalpy
(kW) (kW)
- - 0 - 0.1644

1 -0.0149 -0.0149 -0.0149


0.1495
2 -0.14946 -0.1644 -0.14946 0.0000

3 0.273541 0.1092 0.273541 0.2735

4 -0.00106 0.1081 -0.00106 0.2725

5 -0.21926 -0.1111 -0.21926 0.0532

15
Table 8-3 shows that the pinch temperature is 628.54°C, the minimum hot utility (QH)
is 0.1644 kW, and the minimum cold utility (Q c) is 0.0532 kW. A heat exchanger
network can be designed (Figure 8-2):

Figure 8-2: Heat exchanger network design


The percentage of energy saved can be processed by the below calculations:
Δ H hot =CP Δ T=0.00149 ( 1026.85−623.54 ) =0.6009 kW
(0.6009−0.1643)
×100 %=72.7 %
0.6009
Δ H cold =CP ΔT =0.00114 ( 633.54−226.85 ) =0.4636 kW
0.4636−0.0532
×100 %=88.5 %
0.4636
The percentage of energy saved is 72.7% from the hot utility and 88.5% for the cold
utility.

9. Process Control
In biomass gasification, the most important variable to control is the quality (i.e.
composition) and yield of the syngas produced. To control this, the main parameters
controlled are temperature and pressure.
1. Air flowrate and temperature - Air supplies oxygen for combustion of the
feedstock, affecting the residence time. This means air flow rate controls the
degree of combustion, influencing the gasification temperature. Higher airflow
shortens residence time, but increases temperature. Temperature is
measured by installing a thermocouple in the middle and top of the reacting
vessel. Increasing temperatures increase gas yield of the process, as well as
decrease gas heating value. Additionally, lower solids and tar emissions are
observed.
2. Pressure - Elevated pressures (measured by pressure sensors) serve as a
time-saving step for applications where the gas requires compression
afterwards. However, doing this introduces complexity in feeding the system,
and is very costly (up to four times as much as atmospheric systems (Couto

16
et al., 2013)). This disadvantage may arguably be countered by the higher
efficiency of the system.
3. Gas composition - Measured by passing the product gas through a hot filter,
dew point analyser, rotameter, membrane pump, gas cooler, and gas
analyser(Kumar, Gupta and Viswanadham, 2018).
4. Gasifying agent - The heating value and the hydrogen content of syngas is
affected by the gasifying agent: steam provides higher values than air.
5. Control systems – commercial gasification control systems commonly employ
PLC and SCADA control systems to optimise the gasification process.

10. Liquid-Liquid Extraction


The main PIU operations of extraction, distillation, and crystallization are not
conventionally used in the biomass gasification process. (Wei et al., 2014)
describes a novel idea of using liquid-liquid extraction for the bio-oil produced in
biomass pyrolysis. Complex organic compounds such as alkenes, aldehydes and
alcohols interfere with upgrading of bio-oils, as they cause the formation of coke
(through polymerization and polycondensation reactions). To minimise these
reactions, the oxygen-containing compounds can be extracted from the mixture.

The following is adapted from Wei et al. (2014). For simplicity, the system is
assumed to comprise of three parts: A is the solute (organics), B is the initial
solvent (water), and C is added solvent (hexane).

Table 10-1: Initial conditions of LLE problem from Wei et al. (2014)

A (organic B (water) C (hexane)


compounds)

Feed (L0) 19.09 wt% 80.91 wt% -

Solvent (Vn+1) - - 100 wt%

Extract Target 5 wt% - -

Theoretically, the phase diagram for A and C can then be utilised, along with
equilibrium data, to calculate the mixing point (M) and thus the number of stages
required to meet the extract target of 5 wt% organics. However, due to LLE being
novel in this process, and the complexity of the organic compounds involved, this
data is not easily attainable.

11. Membrane Separation


Membrane treatment systems are suitable for materials with low molecular weight,
that are finely dispersed as solids, or are biological and sensitive. Membranes have
advantages such as low capital and operating costs and the ability to operate at
ambient temperatures. For a given application, membrane suitability differs; the

17
nominal molecular weight cut-off (MWCO) is a parameter used by manufacturers of
how solute molecules are rejected by membranes.
In biomass gasification, membrane treatment systems are used at the end of the
process, aiming to purify the biomass-derived syngas to separate hydrogen. The
mechanisms by which gas separation occurs differs based on the partial pressure
difference (passive) or driving force (active), as well as operating temperatures.
Porous membranes allow convective flow, Knudsen diffusion, and molecular sieving,
whereas dense membranes facilitate solution diffusion(Poudel, Choi and Oh, 2019).
Membrane separation can further be categorised based on membrane material:
polymeric, dense metal, and microporous membranes each have their pros and
cons.
The solute rejection coefficient is defined as:
cp
R=1− ( )
cf
Eq .(11−1)

The MWCO is when R=0.95 and vary from 2,000-100,000 kg/kmol. The general
membrane equation is:
| Δ P|−| Δ π|
J= Eq .(11−2)
( Rm + R c ) μ
Where J is the membrane flux (m3/m2), ∆P is the pressure difference across the
membrane, ∆π is the osmotic pressure difference across the membrane, R m is the
resistance of the membrane, and Rc is the resistance of layers, the filter cake and gel
foulants deposited on the membrane. µ is the viscosity of liquid permeate. Osmotic
pressure is defined as
n
π= RT Eq .(11−3)
Vm

Where n is kmol of solute, V m is the volume of pure solvent water in m 3, associated


with n kmol of solute, R is the gas law constant of 82.057 x 10 -3 m3.atm.kmol-1.K-1.

12. Conclusion
As renewable energy develops, Chemical Engineering students must be able to use
their skills in unfamiliar situations. This coursework shows that the first two years of
the programme allows a student to begin analysing real-world engineering scenarios.
Though they may not have in-depth knowledge of any one area, the foundations
provided by the programme allow the student to navigate literature covering different
aspects of the problem.
Biomass gasification was able to be explored in terms of heat and mass balances,
heat transfer, mass transfer, thermodynamics, process control and process
integration. This report provides insight into the developing industry and assurance
that students are equipped with viable tools for approaching engineering problems.

18
However, more work is required in exploring relationships between modules to lend
the report more continuity, most apparent in sections 10 and 11.

13. Appendix

A Critical Reflection on the Bridging Coursework


As an exchange student, the lockdown effectively served as a premature end to my
time in the UK. Three months sooner than expected, with classes finished and
exams cancelled, I had no reason left to remain. I’d like to say I accepted the logic
amicably, but the reality took a while to sink in: the year, rather than tapering slowly
to a point, was already over. There would be no chance to give proper goodbyes to
the friends I had made there.
As one could expect, this had an impact on my ability to focus on academics (not
mentioning the fourteen-day obligatory quarantine on re-entry to Malaysia). And, with
the pandemic affecting people’s lives around the world, many of my peers reflected a
similar sentiment: what value is there in focussing on exams or coursework?
The decision for exams to be substituted by this Bridging Coursework was welcomed
by most students, including me. Expecting students to dedicate the mental energy to
focus on studying when the global situation shifted daily would be seen as
unempathetic, especially considering international students and the potential turmoil
they face travelling home.
But, conversely, expecting students not to do anything would be unreasonable as
well. For one, we needed something to show for our last academic year. Secondly,
having something for us to work towards provides a sense of structure, preventing a
directionless amalgamation of days. With exams out of the question, this Bridging
Coursework seemed an acceptable compromise.
As with most degree programmes, the first two years builds foundational knowledge.
Though essential, it is easy to compartmentalise modules and not see the
relationships between them; only when the project is sufficiently broad, as in the
Group Design Project, are you required to reach across subjects to link knowledge.
The same applies to this coursework as well.
Given the recommendation to remain within the scope of the briefing, I began with
browsing the literature. Plastics depolymerisation (cracking) seemed dry, and
cryogenic energy storage was outside my realm of interest. Instead, I gravitated
towards energy from waste, as I had some prior knowledge of it. Additionally, it was
a field I aimed to work in, so it seemed wise to use this coursework as an opportunity
to expand my knowledge in it. 
On beginning the coursework, each module offered a different lens to analyse
biomass gasification: PIU analysed how the material is dried before entering the
system; MHM facilitated understanding of transport phenomena for a char particle in

19
the reactor (including how it heats up and diffuses through the system); CET offered
deeper insight into thermodynamic modelling of the reactions. However, it was
quickly made apparent that the literature did not separate cleanly into the forms we
had learnt. Rather, topics were interlinked such that mass transfer would lead to heat
transfer, and that would feed into thermodynamics.
This theme would carry on throughout the coursework: it became clear that modules
separately provided only individual pieces of the solution to a practical Chemical
Engineering problem. An understanding of the underlying relationships between
modules was necessary in order to comprehensively analyse a problem (and thus
effectively solve it). Though I attempted to bridge the gaps between modules, I
consider my efforts only partially successful; a lack of experience in doing similar
exercises has led to this coursework being more apportioned than I intended, with
continuity being difficult to fashion.
In terms of consolidating the past year’s learning, this coursework was only partially
successful. Mainly, much of what we learnt was inapplicable to the scope of the
process: PIU suffered the most from this; there was little need to review processes
such as distillation, extraction, crystallisation, absorption, and membrane separation.
Heat transfer through fins as well could be ignored. The result of this was gaps in my
knowledge.
Exam revision has a stronger case for consolidating learning - by design, a
systematic review of all topics is required when studying for exams. However, even
this has its drawbacks: ‘studying for exams’ usually ensures only short-term retention
of learning. If the goal was long-term retention and application of knowledge to
unfamiliar problems, coursework then is a much better avenue to pursue. From
repeated exposure to a topic, reflecting on problems to be solved, coursework leads
to more assured knowledge. The fact that the scope of knowledge may be narrow is
an acceptable compromise.
In terms of preparing for next year, the value of this exercise is not clear to me.
Though it arguably provides a ‘soft’ taster of the Group Design project, lending us
experience in approaching a project of considerable breadth, the content of the two
projects will likely be different. Moreover, as this exercise doesn’t require us to
review all of last year’s learning, we are left with gaps in our knowledge beginning
Year 3. Of course, one can say that a student with a proper attitude would review the
past year regardless of its utility in the coursework, but I don’t think it’s contentious to
say this applies to a minority. It’s another plus for exams that, even though exam
revision may be for short-term knowledge (i.e. superficial), they force students to do
any revision at all.
As a milestone to graduation, this exercise is useful in two ways: firstly, it indicates
that we’re able to produce work in a pandemic (to employers this should be a sign
that we can remain focussed and reach a goal - though, to be honest, this should be
considered a sign of personal strength, and people should take it as such).

20
Secondly, it expanded my knowledge on biomass gasification so I can point to it in
future interviews and highlight my interest in the energy-from-waste industry.
However, in terms of consolidating a year’s worth of learning, the project’s main goal,
it falls short.
Though that last point sounds negative, it is actually the most important of all. I take
from it that to learn anything (and Chemical Engineering especially) requires active
participation during the course of learning. I learn that for me, neither exams nor
coursework, whilst having their own advantages, serves my goal of really
understanding a topic. To learn anything properly begins with a decision of what
knowledge means to you. 
To that end, and since I want my knowledge to be long-term, strong, and applicable
to different situations, that last point has forced me to assess how I learn altogether. I
understand now what it means to make my learning ‘my own’, and look forward to
how that manifests itself in the future.

14. References
Addenan, M. F. A. (2010) ‘Heat Integration Study of Biomass Gasification Plant for
Hydrogen Production’, International Journal of Heritage Studies.
Antal, M. J. et al. (2000) ‘Biomass gasification in supercritical water’, Industrial and
Engineering Chemistry Research, 39(11), pp. 4040–4053. doi: 10.1021/ie0003436.
Arizaleta, M. L. (2018) ‘Constant-volume Carbonization of Biomass’, C.
Chandra, S. and Boravelli, T. (2016) ‘Design , scale-up , six sigma in processing
different feedstocks in a fixed bed downdraft biomass gasifier’.
Couto, N. et al. (2013) ‘Influence of the biomass gasification processes on the final
composition of syngas’, Energy Procedia. Elsevier B.V., 36, pp. 596–606. doi:
10.1016/j.egypro.2013.07.068.
D., L.-P. et al. (2015) ‘Supercritical Water Gasification of Biomass for Hydrogen
Production: Variable of the Process’, Food and Public Health, 6(3), pp. 92–101. doi:
10.5923/j.fph.20150503.05.
Koufopanos, C. A. et al. (1991) ‘Studies on Kinetics , Thermal and Heat Transfer
Effects’, The Canadian Journal of Chemical Engineering, 69, pp. 907–915.
Kumar, A., Jones, D. D. and Hanna, M. A. (2009) ‘Thermochemical biomass
gasification: A review of the current status of the technology’, Energies, 2(3), pp.
556–581. doi: 10.3390/en20300556.
Kumar, G. S., Gupta, A. and Viswanadham, M. (2018) ‘Design of lab-scale downdraft
gasifier for biomass gasification’, IOP Conference Series: Materials Science and
Engineering, 455(1). doi: 10.1088/1757-899X/455/1/012051.
Molino, A., Chianese, S. and Musmarra, D. (2016) ‘Biomass gasification technology:
The state of the art overview’, Journal of Energy Chemistry. Elsevier B.V., 25(1), pp.
10–25. doi: 10.1016/j.jechem.2015.11.005.
Poudel, J., Choi, J. H. and Oh, S. C. (2019) ‘Process design characteristics of

21
syngas (CO/H 2 ) separation using composite membrane’, Sustainability
(Switzerland), 11(3). doi: 10.3390/su11030703.
Ptasinski, K. J. (2015) Efficiency of Biomass Energy, Efficiency of Biomass Energy.
doi: 10.1002/9781119118169.
Shrestha, R. (2014) ‘Experimental Analysis and Modeling of Biomass Gasification
using a Downdraft Gasifier’. doi: 10.4324/9781315853178.
Sreejith, C. C., Arun, P. and Muraleedharan, C. (2013) ‘Thermochemical analysis of
biomass gasification by Gibbs free energy minimization model - Part: I (Optimization
of Pressure and Temperature)’, International Journal of Green Energy, 10(3), pp.
231–256. doi: 10.1080/15435075.2011.653846.
Wei, Y. et al. (2014) ‘Liquid-liquid extraction of biomass pyrolysis bio-oil’, Energy and
Fuels, 28(2), pp. 1207–1212. doi: 10.1021/ef402490s.

22

You might also like