Sunteți pe pagina 1din 14

EAAW VII Symposium Proceedings 20-21 June 2017

Time Domain Model of a Lead-Acid Cell

D J Harrington*

*BMT Defence Services Ltd, Bath, UK

*Corresponding author. Email: DHarrington@bmtdsl.co.uk / DH530@bath.ac.uk

Synopsis

The lead-acid battery was the first secondary (rechargeable) battery ever to be developed. Invented by French
Physicist Gaston Planté in 1859, its development over the years has led to numerous applications from
smaller automobile starting, lighting and ignition (SLI) batteries up to larger traction batteries for marine
applications. It is important to users of a cell to know how it behaves under different load conditions,
temperatures or a range of other variables. Most lead-acid cell performance data is empiricaland it could be
argued that since empirical data comes from performance measurements of real batteries it is more reliable
than a mathematical model, in which simplifying assumptions have been made in order for it to be tractable.
However, users often have to interpolate between set curves from known data and subsequently they estimate
performance which reduces the accuracy. To improve engineering design, this paper seeks to develop a
reliable, theoretical discharge model of a flooded 8800 Ah lead-acid cell in the time domain in order to
observe how the cell behaves throughout a discharge cycle, theoretically as opposed to experimentally.

Keywords: Energy storage; Electrochemistry; Mathematical modelling.

1. Introduction

In development for over 150 years, lead-acid cells have become ubiquitous in the modern age representing
approximately 70% of the world’s secondary-battery market [1]. Despite recent developments in lithium
ion/lithium polymer cells, advancements are still being made to lead-acid cells due to their numerous benefits. In
an article on modern submarine technology, Dr Randolph Teppner states that Lithium ion cells provide increased
energy density, better voltage stability and increased life-cycle. However the high energy density of the
chemistry in a lithium ion battery results in a need for a rugged cell design with a sophisticated control system to
prevent dangerous operating conditions [2].

Table 1: Advantages and disadvantage of lead-acid cell

Advantages Disadvantages

Low-cost cell (relative to Lithium based cells). Relatively low cycle life (300-500 cycles).

Robust: tolerant to overcharging and can be left Low energy density (30-40 Wh/kg).
on trickle or float charge for extended periods.
Incorrect operation can cause irreversible damage
Good operating temperature (-40°C to 60°C).
(e.g. sulfation of battery electrode).
Voltage varies significantly with charge (~2.1-
Good single cell voltage (>2V).
1.6V).

Available in a variety of designs and


Electrolyte is acidic and liquid.
specifications.

Author’s Biography
Daniel Harrington is a 2nd Year Integrated Mechanical & Electrical Engineering student, studying at
the University of Bath. He is currently undertaking his year in industry at BMT Defence Services in
Bath where he has carried out R&D projects as well as other surface ship related projects within the
naval engineering department.

1
EAAW VII Symposium Proceedings 20-21 June 2017

Table 1 shows the many advantages of the lead-acid cell and why they are so widely used in modern
applications. Despite some disadvantages, the reliable and versatile nature of the lead-acid cell combined with
low cost make it a favoured choice in many applications particularly the maritime defence industry. Lee states
that all Royal Navy submarines past and present use or have used flooded lead-acid cells, and that is still the
only technology qualified for the purpose [3]. This is due to the reliability and vast operating ranges of the cell
which are both essential factors for use at sea, a potentially hazardous environment where cell failure could have
life-threatening consequences. It is not only important that the cell does not fail but also that the crew can
predict reliably how the cell is going to operate in certain conditions.

In this paper, a mathematical model of a lead-acid battery is presented. The model has been developed from first
principles and uses simplifying assumptions to demonstrate the different electrochemical processes at work
inside the cell at any one time. Example results will be presented and compared to empirical data obtained from
a proprietary battery data sheet. The paper concludes with a discussion of the results highlighting any infidelities
in the model, as well as the assumptions made.

2. Fundamentals of a lead-acid cell

2.1 Structure

The simple structure of a lead-acid cell comprises two electrodes immersed in an electrolyte solution of sulfuric
acid (H2SO4) – the anode is made of lead (Pb) and the cathode of lead dioxide (PbO2). Figure 1 shows a cutaway
diagram of a lead-acid cell.

Vent plugs
Cell pillars and connectors

Porous Lead Anode (-)


PbO2 Cathode (+)

Plastic container
Electrolyte: Sulfuric Acid solution

Figure 1- A cutaway diagram of a lead-acid cell [4].

The porous lead anode contains many cavities to maximise the surface area of lead in contact with the
electrolyte, thus maximising the Electromotive Force (EMF) produced at the terminals. While there are
numerous variations of the lead-acid cell, such as valve regulated cells or cells that have lead grids coated with a
paste of lead oxides, this paper will consider a flooded lead-acid cell (electrodes in an excess of electrolyte).

2.2 Chemistry

Electrochemical cells produce an EMF across their terminals due to oxidation and reduction reactions occurring
at the electrodes. The lead-acid cell combines Pb and PbO2 electrodes in sulfuric acid to produce an EMF. The
cell utilises the chemical reactions of lead oxidation (loss of electrons) and lead dioxide reduction (gain of
electrons) to cause a flow of charge and hence produce an EMF across the terminals.

2
EAAW VII Symposium Proceedings 20-21 June 2017

The electrode equations are as follows. At the lead electrode (negative plate):

        (1)

At the lead dioxide electrode (positive plate):

              (2)

Equation 1 shows that the lead reacts with aqueous sulfate ions, produced by the sulfuric acid solution, to
produce lead sulfate (  and two electrons. Equation 2 shows that the two electrons produced in the first
reaction reduce the lead in PbO2 allowing lead sulfate and water to form. Equation 2 also shows H2SO4 being
converted to water through reaction. This means that the amount of acid in the electrolyte reduces with cell
discharge; an important relationship used for the model.

The result of each of these chemical reactions is that each electrode has potential measured in relation to a
standard hydrogen electrode. The lead electrode has a negative potential and the lead dioxide has a positive
potential and this consequently causes a potential difference between them, giving the cell an EMF which can be
calculated by Equation 3 below [5].

      (3)

The variation of the EMF forms a large part of the model and will be discussed later in the paper.

3. Modelling approach

The first key aspect of designing the model was establishing known cell relationships for which equations could
be derived and subsequently implemented in Matlab®. Firstly, it was known that the Relative Density (RD) of
the electrolyte reduced with the State of Charge (SOC) of the cell, from approximately 1.28 at full charge to
approximately 1.08 at complete discharge [6]. This change in RD represents a change in concentration of the
electrolyte as sulfuric acid has a much higher density than that of water (pure sulfuric acid has an RD of around
1.84 and water has an RD of 1); at full charge there is a higher concentration of acid giving it a higher RD, but
as the acid reacts through discharge the solution becomes more aqueous causing the RD to approach 1.

The voltage of a lead-acid cell reduces with SOC (from around 2.1V to 1.6V) with a non-linear relationship
showing that the voltage is proportional to the concentration of the electrolyte solution in which the electrodes
are placed. The equation that relates these variables is known as the Nernst equation, formulated by German
physical chemist Walther Nernst [7]. The specific Nernst equation for the lead-acid cell can be seen below (for
derivation see [8]).

 
        )

Where:
Ecell = EMF of cell
E0 = Standard cell potential
R = Molar gas constant
T = Temperature in Kelvin
n = Number of moles of electrons involved in reactions
F = Faraday’s constant
aH2SO4 = activity of H2SO4

3
EAAW VII Symposium Proceedings 20-21 June 2017



In chemical thermodynamics, activity is the measure of the effective concentration of a species in the mixture
under non-ideal conditions. It is used to determine the chemical potential for a real solution rather than an ideal
one. Equation 5 shows how to calculate the activity of the electrolyte solution [8].


        

Where γm is the mean activity coefficient and C is the molal concentration in mol/kg. Molal concentration is
defined by the number of moles of solute (acid) per kilogram of solvent (water). The mean activity coefficient is
generally temperature and concentration dependent and is determined experimentally.

Once this relationship was recognised, the equations that governed the change in the concentration of the acid
with time were derived. To simplify this task a basic model of the cell was established which can be seen in
Figure 2.

Semi-permeable boundary
Bulk Reservoir of Acid

Active acid

Lead Electrode

Holes in Porous lead electrode

Figure 2 – Simplified model of Lead-Acid Cell

Figure 2 shows the simplified model which has been broken down into two separate volumes separated by a
semi permeable membrane (porous boundary of electrode): the bulk volume of acid which is the large reservoir
of acid not in contact with the electrode and the active volume which is the acid in contact with the electrode
either on its surface or in the holes. The model only contains the lead anode as this is where the acid reacts and
its porous nature allows for a diffusion model to be set up.

The aim of this simplified model was to establish how the mass of acid in the active volume and bulk volume
changed with time; this meant deriving equations that would give the mass of acid in each volume as a function
of time. The derivation from first principles of these equations can be seen in Appendix A. For the purpose of
this paper subscript 1 denotes the bulk volume, subscript 2 denotes the active volume and subscript 0 denotes an
initial condition.


4
EAAW VII Symposium Proceedings 20-21 June 2017

  (6)
            
     
     

  (7)
            
     
     

Where:

M = Mass of Acid

V = Volume

σ = Permeability coefficient (value between 0-1, where 0 is completely permeable and 1 is a solid membrane)

t = Time

As the concentration of acid is directly proportional to the mass, the equations made it possible to effectively
model the variation in concentration throughout the discharge cycle and subsequently model the variation in
EMF.

Another aspect of the cell that had to be modelled was the heat transfer throughout the cell. This was due to the
fact that the internal resistance of the cell changes with temperature, as does the electrical conductivity of the
electrolyte. This in turn affects the EMF due to the T term in the Nernst equation, so modelling the heat transfer
had improved accuracy over using a fixed temperature. There is an increase in internal temperature of the lead
electrolyte due to I2r (lower case r denotes internal resistance) resistive heating; this heat is then transferred to
the active acid solution and this heat is in turn transferred to the bulk solution.

The heat transfer was modelled using the same equations derived from the concentration model as the variables
map directly. The variable mapping can be seen in Table 2.

Table 2: Variable comparison of concentration model and thermal model

Concentration Model Heat transfer Equivalent


Symbol Symbol
Variable Variable

Volume V Heat Capacity C

Mass of acid M Heat energy Q

Flow of acid dM/dt Heat Flow dQ/dt


Driving Head
H Temperature difference ΔT
(concentration difference)
Permeability σ Thermal conductivity Z

Once these relationships were derived, it was then possible to implement the equations in Matlab® in order to
test the model and obtain visual outputs of the results. To represent the code, Figure 3 shows a flow diagram
which highlights the key processes and calculations carried out in the code in order to give the outputs.

5
EAAW VII Symposium Proceedings 20-21 June 2017

0  
   ,
0 !166552
0  ""#
0  !#&!
0 #!$"
$$" 0 &!!#

0    


       2
0 &#$''&!&""!!$##- ##!&#
0 "# &"$#($'&/(
&#)#
#!$" "-
0 
(#!$&"#&#( !
!"# &$- !
 /
0 
(#'* 
 
"# 0  &"$#&!"'!"#!! &$"-

0 !'#'!""#!"!!*" #3-
0 !!*"&"##!#!$""-

"&#"

Figure 3 – High level flow chart highlighting the key stages in coding of lead-acid cell model.

To obtain a solution, a numerical solver was used to compute the analytic solution to the initial value problem.
As a positive current is being drawn from the cell, the total charge reduces by the current multiplied by the time-
step, which in turn causes a proportional drop in the mass of acid in the active volume. This is equivalent to the
acid being transformed into water through chemical reaction in a real cell. The first order differential equations
then provide an analytic solution to the acid equalisation process within the time-step to provide new masses of
acid in the active and bulk volumes. This gives new concentrations and therefore provides an analytical solution
to the Nernst equation in each time-step.

The method accounts for reaction and equalisation of the acid in the same time-step as two consecutive
processes and uses what is essentially a simple Euler method combined with first order differential equations to
solve the problem. By using a short time-step of 1s and an analytical solution from the differential equations, the
inaccuracy of the method is reduced to an acceptable value consistent with the accuracy of the original data.
Furthermore less computing power is required than for a higher order method such as the Runge-Kutta 4th Order
method.

6
EAAW VII Symposium Proceedings 20-21 June 2017



4. Results

This section of this paper presents the variation of different parameters throughout discharge cycles of the cell.
The model was run under different conditions to see the variation in performance.

The first test run on the model was to see how the cell voltage varied for time with different discharge rates.
This was achieved by running the model for a number of different load currents. The plots from the model can
be seen in Figure 4a, whilst Figure 4b shows empirical results from an 8800 Ah cell with the same load currents
(Figure 4a shows some empirical data points overlaid for easy comparison). Due to the sensitive nature of the
original data, the current values have been removed.

x &!!# !#

x &!!#!#

1x&!!#!#

x &!!#!#

x &!!#!#

x &!!#!#
#.

.!"
Figure 4a and 4b – Comparison of model discharge curves to real data for a range of load currents, showing
variation in voltage with respect to time


7
EAAW VII Symposium Proceedings 20-21 June 2017



Another output from the model that can be observed is the change in concentration over one cycle; this was the
main factor in changing EMF so was an important variable to consider. To show the variation in concentration,
the model was run for a 32000s simulation: 8000s at a high load current, 8000s with no current, 8000s with a
medium load current and another 8000s of no current. Figure 5 shows the load current profile for the simulation.
The resulting variation of concentration of acid in the lead (active) volume, bulk volume and the level
concentration with time can be seen in Figure 6. Figures 7 subsequently show the variation in cell EMF over the
period modelled.

Figure 5 – Variation of load current with respect to time

Figure 6 – Variation in concentration of acid in the active and bulk volumes with respect to time (the ‘Level
Concentration’ is the total volume of acid in the cell is divided by the total volume)


8
EAAW VII Symposium Proceedings 20-21 June 2017



Figure 7 -Variation of cell EMF with respect to time

Another variation occurring in the cell is the heat transfer. The lead electrodes heat up due to the internal
resistance of the cell and this in turn is transferred to the acid via conduction. Figure 8 shows the variation in
temperature of the solid lead, lead (active) solution and bulk solution over a typical discharge cycle from fully
charged to fully discharged with a load current of 525 A.

Figure 8 – Variation of temperature with respect to time


9
EAAW VII Symposium Proceedings 20-21 June 2017



5. Discussion

Figure 4a & 4b form the most important evidence for the validity of the model designed. It is observed that the
model gave intuitive results when compared to empirical data, with the same characteristic shape of curve. This
shows that that the basis for the variation in EMF caused by the changing concentration of the electrolyte is
valid for the model. Overall the accuracy of the model is good with all plots having less than 7% inaccuracy
when compared to the empirical data. The least accurate plot was the 3.5x current plot which had a mean
difference of 6.3% and the most accurate plot was the 0.3x plot which had a mean difference of 0.86%. As the
model contains simplifications of the highly complex electrochemical and physical processes in the cell, this
level of accuracy supports the validity of the model.

Figure 6 shows that when a load current is applied there is an initial steep drop in the concentration of acid in
the active volume. This is because the active volume is the location of the chemical reaction where acid reacts to
produce an EMF. After an initial drop the rate of change reduces as acid from the bulk volume diffuses into the
active volume to replenish the lost acid. After some time the rate of loss of acid from the active volume is equal
to the rate of diffusion of acid which causes the two lines to have the same gradient. The ‘Level Concentration’
is the concentration obtained if the total volume of acid in the cell is divided by the total volume. As there is a
constant current being applied the level concentration has a constant negative gradient.

Once the load current is removed, Figure 6 shows an equalisation process at 8000s. As there is no chemical
reaction occurring once there is no current, the concentration of the acid in the active volume does not reduce.
However, the cell does not remain at a steady state as there still remains a difference in concentration between
the bulk and active volumes. This causes an equalisation process to occur as the concentrations in the two
volumes tend towards a new steady state value where the concentrations are equal. The change in the active
volume is much more noticeable due to it having a significantly smaller volume.

At 16000s a new lower current is applied. This causes a softer initial drop to the active volume concentration as
less acid is being removed per unit time; this is also the reason for the smaller rate of change of concentration.
Figure 6 also shows a smaller separation between the parallel bulk and active concentration lines – this is due to
less time passing before the rate of loss of acid from the active volume equals the rate of diffusion. The same
equalisation can be seen at 24000s however as the concentration head is smaller, less time is needed for
equalisation.

Figure 7 shows the variation in EMF for the same discharge cycle. There is an initial drop in EMF when the
current is applied due to an initial drop in concentration of the active acid. However, the concentration of the
active acid does not drop low enough in the cycle for the significant non-linearity of the EMF curve to be
observed, meaning that the gradient remains relatively constant. Once the current is switched off and the
equalisation occurs the EMF can be seen to increase slightly; this is due to the concentration of acid in the active
volume increasing slightly when the equalisation process occurs. As the EMF is directly related to the
concentration of active acid, there is a subsequent increase in EMF.

The same process can be seen when the lower current is applied however this time the rate of change of EMF is
reduced as there is less acid reacting per unit time due to the lower current. The second equalisation causes a
much smaller increase in EMF as there is a smaller increase of the active volume concentration to the new
steady state value.

Figure 8 shows the variation in temperature of the constituent parts of the cell over a discharge cycle. As the
lead electrode is the part of the cell that is being heated directly by I2r heating, an initial increase in temperature
can be seen as heat energy enters the mass directly. As the heat is dissipated into the active acid volume the rate
of change reduces. As the electrolyte has a much higher heat capacity compared to the solid lead the rate of heat
transfer from the solid lead to the active solution is slow, therefore it takes some time before a change in
temperature can be detected. As the bulk volume is large compared to the active volume, an increase in the heat
energy of the active volume has almost no effect on the overall temperature of the bulk volume, hence the
temperature of the bulk volume remains almost constant.


10
EAAW VII Symposium Proceedings 20-21 June 2017



This model contains many assumptions and simplifications of the physical and electrical behaviour of a real cell.
This was necessary due to the extremely complex electrochemical processes occurring within the cell. The
model aims to effectively represent the processes in the cell and give the best possible outputs whilst adhering to
the computing power available.

Two key assumptions made were the values of the electrode permeability coefficient and the ratio of volumes of
the bulk to the active acid. These values were required to accurately model the cell; however it was not possible
without experimentation to obtain values for these variables. Values were therefore estimated, and subsequently
iterated, to achieve the most accurate outputs. It was estimated that the active volume was 8% of the bulk
volume (5% was the initial rough estimate and the value was then iterated) and the permeability coefficient of
the boundary was set at 0.015. Whilst these are estimates, the real value of each is cell dependant and varies
with design. Moreover these values gave accurate outputs from the model with an error of less than 7%;
therefore they can be taken as adequate for the purposes of this paper.

Another assumption was the manner of heat transfer in the cell. This was modelled as firstly heat energy going
directly into the lead electrode then in a two stage conduction from the lead electrode to the active volume, and
then from the active volume to the bulk volume. In a real cell there would also be convection and mixing of the
two volumes which would cause heat transfer but these processes are highly complex and therefore would be
hard to model accurately. Despite the inaccuracies, the changing temperature of the active acid did not have a

large effect on the EMF produced as the term in the Nernst equation had only a 9.7% variation over the

temperature range in the model.

Finally an important relationship in the model was that between molal concentration and γm. The relationship
was defined in the model from experimental data [5] using interpolation. This data is recorded at a constant
temperature of 298.15 K and the model uses a changing electrolyte temperature, therefore this will cause some
errors. Nevertheless, the temperature change was not significant and as γm cannot be recalculated at each time
step the assumption was necessary to enable modelling of the system.

6. Conclusion

In conclusion, the model gave accurate and intuitive results when compared to empirical data. This validates the
model as it used simplified processes of the action within the cell yet still gave the same characteristic outputs.
Nevertheless the overall simplification of the electrochemical processes meant that results did not match up
perfectly, particularly at high and low load currents. The concentration profile shows that the first order
differential equations derived are correct as they follow the right shape for the load conditions of the cell (i.e.
equalisation at 0 current, constant gradients with a load current). The subsequent EMF curve supports the
proportional relationship between concentration of the electrolyte and the cell EMF. The temperature profile
showed that the equations determining the transfer of heat gave intuitive outputs despite the simplifications;
however this part of the model had little effect on the results and therefore could have been neglected as the
temperature change of a lead-acid cell remains within a relatively small range.

With some improvement, this model could be used as an on-board tool to predict battery performance. By
inputting a known load cycle into the model, the user could observe predicted model performance and use this to
adjust the actual loading in order to meet a set of desired behaviour.

Acknowledgements

The guidance, advice and support of Prof Chris Hodge OBE FREng are acknowledged with thanks.

The kind permission and resources granted to the author by BMT are acknowledged with thanks.

All findings, ideas, opinions and errors herein are those of the author and are not necessarily those of BMT
Defence Services Limited.


11
EAAW VII Symposium Proceedings 20-21 June 2017



References
1. Jung, J.J, 2015. Lead-Acid Battery Technologies: Fundamentals, Materials and Applications. 1st ed.:
CRC Press.
2. Teppner, R, 2013. Modern Submarine Technology. Marine Engineers Review (MER), Issue
July/August 2013, Page 27.
3. Lee, J.M, Faulkner, D & Rowlinson, P, 2010. Increasing the battery voltage on a nuclear submarine.
Proceedings of International Naval Engineering Conference (INEC). Portsmouth, 11-13th May 2010.
4. Yoder, J.A, 1996. DOE Primer on Lead Acid Storage Batteries. 1st ed. Page 25: U.S. Department of
Energy. [DOE-HDBK-1084-95]
5. Pavlov, D.P, 2011. Lead-acid Batteries Science and Technology: a handbook of lead-acid battery
technology and its influence on the product. 1st ed. Pages 30-33: Amsterdam: Elsevier Science Ltd.
2011.
6. Staples, B.R, 1981. Activity and Osmotic Coefficients of Aqueous Sulfuric Acid at 298.15 K. J. Phys.
Chem. Ref. Data, Vol. 10, No. 3, Pages 779-798. Available at: https://srd.nist.gov/JPCRD/jpcrd186.pdf
[Accessed 18 September 2016].
7. Orna, Mary Virginia; Stock, John (1989). Electrochemistry, past and present. Columbus, Ohio:
American Chemical Society.
8. Berera G.P, 2006. 3.014 Materials Laboratory Module. Lead-Acid Storage Cell, [Online]. 1,4.
Available at: http://ocw.mit.edu/courses/materials-science-and-engineering/3-014-materials-laboratory-
fall-2006/labs/w3_b1.pdf. [Accessed 19 September 2016].

APPENDIX A


Figure 9 - Schematic Representation of Cell Diffusion

 

 The total volume of the cell available for acid

 The bulk volume of the reservoir of acid

 The volume of the acid contained in the porous lead (active region)

 The concentration of the acid if fully equalised between the two volumes

 The concentration of the bulk volume acid

 The concentration of the active acid

 The total mass of the acid

 The mass of the acid in the bulk volume acid

 The mass of the acid in the active volume


12
EAAW VII Symposium Proceedings 20-21 June 2017



 The initial mass at time zero of the acid in the bulk volume acid

 The initial mass at time zero of the acid in the active acid

H The concentration head (concentration difference)

 The permeability of the membrane (given by a factor between 0-1, where 0 is no boundary and 1 is
solid wall)

Relationships
     Volume of cell is constant – no loss of acid

     Mass is constant (conservation of mass)


  Concentration is mass per unit volume



 


   
 
 


       The rate of change of mass is given by

permeability factor multiplied by the
concentration head (difference in concentration).

(Note: a positive flow of acid from the bulk volume to the porous lead reduces the mass in the bulk volume
hence the negative sign of the RHS.)

Differential Equation

      


  
  
  

          

      

    


   Substitute in to remove M2 term
  

    


   
   

    


   
  

     


   
   

  
      Rearrange to get M1 terms on the left hand side
  
 

    
     
   
    Integrate both sides


 


13
EAAW VII Symposium Proceedings 20-21 June 2017



     


      Combine LHS integral to a single fraction
   

  
     
      
   
 

    


     

     Exponentiate both sides to remove log
  
 

 
       

       Rearrange & simplify to isolate M1
   

 
     

         
   

 
        

     (8)
   

      

 
      

            Substitute in (8) to get equation
   
for M2

 
                  

    
   

 
         

     Rearrange & simplify
   

 
        

     (9)
   

Note:

Adding the expressions for  

Gives:

 
      
   

       


 

       


 

      


 

   


 

  

Which is, as expected, the total mass and constant.


14

S-ar putea să vă placă și