Download as pdf or txt
Download as pdf or txt
You are on page 1of 30

Accepted Manuscript

Title: Dynamic optimization of beer fermentation: sensitivity


analysis of attainable performance vs. product flavor
constraints

Authors: Alistair D. Rodman, Dimitrios I. Gerogiorgis

PII: S0098-1354(17)30270-3
DOI: http://dx.doi.org/doi:10.1016/j.compchemeng.2017.06.024
Reference: CACE 5855

To appear in: Computers and Chemical Engineering

Received date: 11-3-2017


Revised date: 20-6-2017
Accepted date: 23-6-2017

Please cite this article as: Rodman, Alistair D., & Gerogiorgis, Dimitrios I.,
Dynamic optimization of beer fermentation: sensitivity analysis of attainable
performance vs.product flavor constraints.Computers and Chemical Engineering
http://dx.doi.org/10.1016/j.compchemeng.2017.06.024

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Dynamic optimization of beer fermentation: sensitivity analysis of
attainable performance vs. product flavor constraints

Alistair D. Rodman, Dimitrios I. Gerogiorgis*

School of Engineering, University of Edinburgh, The King’s Buildings, Edinburgh, EH9 3FB, UK

*
Corresponding author: D.Gerogiorgis@ed.ac.uk (+44 131 651 7072)

HIGHLIGHTS
 Explicit multi-objective dynamic optimization of beer fermentation is pursued for the
first time.
 Ethanol production maximization and batch duration minimization are simultaneously
pursued.
 The strong influence of flavour (ethyl acetate, diacetyl) constraints is quantitatively
monitored.
 Initialisation (temperature profile) and time discretization affect solution quality
considerably.
 Novel solutions of strong industrial promise are reported; trade-off visualisations are
illustrated.

ABSTRACT
The declining alcohol industry in the UK and the concurrent surge in supply and
variety of beer products has created extremely competitive environment for
breweries, many of which are pursuing the benefits of process intensification and
optimisation. To gain insight into the brewing process, an investigation into the
influence of by-product threshold levels on obtainable fermentation performance
has been performed, by computing optimal operating temperature profiles for a
range of constraint levels on by-product concentrations in the final product. The
DynOpt software package has been used, converting the continuous control vector

1
optimisation problem into nonlinear programming (NLP) form via collocation on
finite elements, which has then been solved with an interior point algorithm. This
has been performed for increasing levels of time discretisation, by means of a
range of initialising solution profiles, for a wide spectrum of imposed by-product
flavour constraints. Each by-product flavour threshold affects process performance
in a unique way. Results indicate that the maximum allowable diacetyl
concentration in the final product has very strong influence on batch duration, with
lower limits requiring considerably longer batches. The maximum allowable ethyl
acetate concentration is shown to dictate the attainable ethanol concentration, and
lower limits adversely affect the desired high alcohol content in the final product.

Keywords: Beer fermentation; dynamic optimization; multi-objective

optimization; orthogonal collocation on finite elements; sensitivity

analysis; flavour constraints

NOMENCLATURE
Roman symbols

𝑊𝐸 Objective ethanol weight (%)


𝑊𝑡 Objective time weight (%)
𝑋𝐴 Active biomass concentration (g L-1)
𝑋𝐷 Dead biomass concentration (g L-1)
𝑋𝐿 Latent biomass concentration (g L-1)
𝑌𝐸𝐴 Ethyl acetate production stoichiometric factor (g L-1)
𝑘𝑒 Ethanol affinity constant (g L-1)
𝑘𝑠 Sugar affinity constant (g L-1)
𝑘𝑥 Biomass affinity constant (g L-1)
DY Diacetyl (-)
EA Ethyl Acetate (-)
EtOH Ethanol (-)
g Inequality constraint
ℎ Equality constraint
K Number of collocation points
N Number of elemets in time horizon

2
S Sugar (-)
T Fermenter temperature (K)
𝑓 Fermentation inhibition factor (g L-1)
𝑡 Time (h)
𝑡f Batch time (h)
𝑢 Model control
𝑥 Model state

Greek symbols

∆𝑡𝑖 Length of element i


𝜇𝐴𝐵 Diacetyl consumption rate (g-1 h-1 L)
𝜇𝐷𝑇 Specific cell death rate (h-1)
𝜇𝐷𝑌 Diacetyl growth rate (g-1 h-1 L)
𝜇𝐸 Ethanol production rate (h-1)
𝜇𝐿 Specific cell activation rate (h-1)
𝜇𝑆 Sugar consumption rate (h-1)
𝜇𝑆𝐷 Specific dead cell settling rate (h-1)
𝜇𝑥 Specific cell growth rate (h-1)
𝛺 State approximation polynomial
𝜑 Objective function (-)
𝜓 Control approximation polynomial

Subscripts and operators

(̃) Normalised parameter (-)


( )0 Initial condition (-)
( )𝐿 Lower bound (-)
( )𝑈 Upper bound (-)
( )𝑖 Property in element i (-)
( )𝑗 Property at collocation point j (-)

3
1. INTRODUCTION
Determining how a modern industrial production process shall be operated typically involves
mathematical optimisation in some form. Often this will include an optimal control problem, where a
system of state variables [𝑥] are influenced by an externally manipulatable control variable, 𝑢, so the
optimal control vector 𝑢(t) is sought to maximise an objective, 𝜑, here considering only a terminal
payoff (Biegler, 2010; Biegler et al., 2012):

min 𝜑(𝑥(𝑡𝑓 ), 𝑡𝑓 ) (1)


𝑢(𝑡),𝑡𝑓

𝑑𝑥(𝑡)
s.t = 𝑓(𝑥(𝑡), 𝑢(𝑡)), 𝑥(𝑡0 ) = 𝑥0 (2)
𝑑𝑡

ℎ(𝑥(𝑡), 𝑢(𝑡)) = 0, 𝑔(𝑥(𝑡), 𝑢(𝑡)) ≤ 0 (3)

ℎ𝑓 (𝑥(𝑡)) = 0 𝑔𝑓 (𝑥(𝑡𝑓 )) ≤ 0 (4)

𝑢(𝑡)𝐿 ≤ 𝑢(𝑡) ≤ 𝑢(𝑡)𝑈 , 𝑥(𝑡)𝐿 ≤ 𝑥(𝑡) ≤ 𝑥(𝑡)𝑈 (5)

The ordinary differential equations (ODEs) which dictate the state trajectories (Eq. 2) are influenced
at any time by the current control (u) value, while Eq. 3 represents equality and inequality constraints
across the entire time horizon, 𝑡 ∈ [𝑡0 , 𝑡𝑓 ], with terminal constraints given by Eq. 4. Lastly the state
and control boundaries are constrained within permissible bounds by Eq. 5.

An investigation into the beer manufacturing industry in the UK has been performed to determine if a
strong incentive for process intensification and optimisation exists. The alcohol industry as a whole
has been in decline in recent years within the UK as shown in Fig. 1, where annual litres of pure
alcohol per capita is the metric used to normalise for beverages of differing alcoholic strength. This a
result of several factors: people are drinking from a later age and regular drinkers are turning away
from high strength products, towards more costly and lower strength drinks, such as craft beer.

Beer is however one of the few exceptions from the trend of a declining sector. The growing market
share fuelled by recent increased demand for high value craft beer products produced on a small scale
has led to the beer industry growing both in terms of production volume and market value. 1% year on
year growth is predicted over the next 3 years, with the annual production volume in the UK expected
to exceed 4.6 billion litres by 2019, compared to 4.2 billion in 2015. Fig. 2 shows the number of
breweries in operation over the last 6 years in the UK: it is evident that there is very steady increase
which is predicted to continue moving forward.
Fig. 3 depicts the UK’s alcohol consumption in context vs the rest of Europe. While Scots may have a
reputation of being heavy drinkers it is evident that while their per capita consumption is above the
average for the rest of the UK, it is still a very typical value within Europe.

4
The result of the declining alcohol industry and the surge in supply of beer products has created an
extremely competitive environment for producers, many of whom must look towards process
intensification if they are to remain profitable, forming the motivation for this study.

Within the beer production process the fermentation stage is generally the system bottle neck, with
batch times in excess of one week not uncommon. Fermentation progression depends on many
variables (Rodman and Gerogiorgis, 2016b), however progression is dominated by the influence of
the temperature of the involved substrates. As such, it is necessary to determine the temperature
manipulation profile capable of steering the process to competition in an optimal manner.

Approaches to process optimisation fall under three areas (Bonvin, 1988):

 off-line optimisation (open loop optimal control)


 run-to-run optimisation
 on-line optimisation

This study is concerned with the former: determining solutions to the off-line optimisation problem to
provide optimal open loop trajectories for the manipulated and state variables. These profiles are
computed once, off-line, thus feedback elements are not included, and rather an ideal recipe for
optimal production is produced. This approach is limited in usefulness as in the presence of
disturbances these trajectories lose their optimal character (Canto et al., 2000), however on-line
optimisation is not practical: online concentration readings are extremely cumbersome to monitor in
many cases. Rather many medium scale breweries elect to take a sample once the prescribed
temperature trajectory has completed and determine the residual sugar content based on the product
density (a surrogate measurement for total sugar content) via the Plato scale, to confirm if the batch
has completed fermentation as expected and desired. This convention renders any attempt to
incorporate an online control loop for control of state (concentration) trajectory control non-applicable
to this particular problem and is the reason why our study is focused on off-line optimisation. A beer
brand or line instead typically has a proprietary temperature manipulation profile (recipe) used for
every batch, to ensure product consistency, which fits the scope of this work.

2. PROCESS DESCRIPTION
2.1 Beer Fermentation

Fermentation is an essential step in the manufacture of alcoholic beverages, responsible for the
characteristic taste of the final product and its alcohol content (Rodman and Gerogiorgis, 2016c).
Upstream processing produces a sugar rich intermediate (wort) from a feedstock starch source (most
typically malted barley). Once cooled to an appropriate initial temperature the wort enters stainless
steel vessels along with yeast, allowing fermentation to commence. The primary chemical reaction

5
pathway is the conversion of sugars into ethanol and carbon dioxide, which is coupled with biomass
(yeast) growth and heat generation from the exothermic reaction. Concurrently, a range of species are
formed at low concentrations by a multitude of side reactions, many of which may impact product
flavour above threshold concentrations. Fermentation is completed once all consumable sugars have
been converted by the yeast into alcohol, following which the solution leaves the fermenter for
subsequent downstream processing prior to sale and consumption.

2.2 Fermentation modelling


Several mathematical models for the beer fermentation process have been published (Gee and
Ramirez, 1988; de Andres-Toro, 1998, Trelea et al., 2001). Models are reduced order, considering
only the key species present due to system complexity (200+ species, Vanderhaegen et al., 2006)
rendering exhaustive modelling extremely cumbersome: in fact to date many of the specific chemical
interactions in the fermentation process are not understood. The kinetic model of beer fermentation by
de Andrés-Toro et al. (1998) has been selected for study due to its direct applicability to the industrial
process:

 Published parameters are derived from a very large array of experiments, resulting in a wide
temperature range (8–24 ºC) which ensures high fidelity and applicability.
 The model includes all prominent by-products which degrade beer product quality in terms of
taste and aroma, rendering the model valuable for assessing performance.
 Predicted profiles indicate the highest fidelity with experimental and pilot-plant data in
comparison to other models, due to successful validation against over 200 fermentations.

The model considers 7 states (Eqs. 6-12), with trajectories governed by temperature dependant
production and consumption factors (Eqs. 13-16). The model structure takes the form shown in Fig. 4.
Yeast cells transition from latent to active to dead over time, with only active cells able to promote
fermentation (conversion of sugar to ethanol). A more detailed description of the model can be found
in its original publication (de Andrés-Toro, 1998), along with the constants for the Arrhenius
relationship governing the parameters temperature dependence, as computed from industrial scale
fermentation data. The exception is that the rates for diacetyl are taken from a later publication, shown
in Table 1 along with the initial state concentrations used for simulation.

𝑑[𝑋𝐴 ](𝑡)
= 𝜇𝑥 (𝑡, 𝑇) ∙ [𝑋𝐴 ](𝑡) − 𝜇𝐷𝑇 (𝑡, 𝑇) ∙ [𝑋𝐴 ](𝑡) + 𝜇𝐿 (𝑡, 𝑇) ∙ [𝑋𝐿 ](𝑡) (6)
𝑑𝑡

𝑑[𝑋𝐷 ](𝑡)
= −𝜇𝑆𝐷 (𝑡, 𝑇) ∙ [𝑋𝐷 ](𝑡) + 𝜇𝐷𝑇 (𝑡, 𝑇) ∙ [𝑋𝐴 ](𝑡) (7)
𝑑𝑡

𝑑[𝑆](𝑡)
= −𝜇𝑆 (𝑡, 𝑇) ∙ [𝑋𝐴 ](𝑡) (8)
𝑑𝑡
𝑑[𝐸𝑡𝑂𝐻](𝑡)
= 𝑓(𝑡) ∙ 𝜇𝑒 (𝑡, 𝑇) ∙ [𝑋𝐴 ](𝑡) (9)
𝑑𝑡
𝑑[𝐸𝐴](𝑡)
= 𝑌𝐸𝐴 (𝑇) ∙ 𝜇𝑥 (𝑡, 𝑇) ∙ [𝑋𝐴 ](𝑡) (10)
𝑑𝑡

6
𝑑[𝐷𝑌](𝑡)
= 𝜇𝐷𝑌 ∙ [𝑆](𝑡) ∙ [𝑋𝐴 ](𝑡) − 𝜇𝐴𝐵 ∙ [𝐷𝑌](𝑡) ∙ [𝐸𝑡𝑂𝐻](𝑡) (11)
𝑑𝑡

𝑑[𝑋𝐿 ](𝑡)
= −𝜇𝐿 (𝑡, 𝑇) ∙ [𝑋𝐿 ](𝑡) (12)
𝑑𝑡

𝜇𝑥0 (𝑇)∙[𝑆](𝑡)
𝜇𝑥 (𝑡, 𝑇) = (13)
0.5∙[𝑆]0 +[𝐸𝑡𝑂𝐻](𝑡)

𝜇𝑆𝐷0 (𝑇)∙0.5∙[𝑆]0
𝜇𝑆𝐷 (𝑡, 𝑇) = (14)
0.5∙[𝑆]0 +[𝐸𝑡𝑂𝐻](𝑡)

𝜇𝑠0 (𝑇)∙[𝑆](𝑡)
𝜇𝑠 (𝑡, 𝑇) = (15)
𝑘𝑠 (𝑇)+[𝑆](𝑡)

𝜇𝑒0 (𝑇)∙[𝑆](𝑡)
𝜇𝑒 (𝑡, 𝑇) = (16)
𝑘𝑒 (𝑇)+[𝑆](𝑡)

[𝐸𝑡𝑂𝐻](𝑡)
𝑓(𝑡) = 1 − (17)
0.5∙[𝑆]0

The model considers two by-product species alongside the primary reaction pathway: ethyl acetate
(Eq. 12) and diacetyl compounds (Eq. 11). Diacetyl (2,3-butanedione) has a pungent butter-like aroma
(Izquierdo-Ferrero et al., 1997), while ethyl acetate is often used as an indicator of all esters present,
and is described as having the odour of nail varnish remover (Hanke, S. et al., 2010). Different beer
products may contain differing levels of these species, since the specific flavour profile will influence
the levels above which these by-products will degrade the flavour. This study has looked into the
influence which imposed limits on these compounds concentrations in the final beer product have on
the attainable fermentation performance and efficacy, by considering a range of realistic thresholds of
each.
The primary source of potential modelling error lies in the model parameterization and the validity of
these parameters for representing any specific brewery process. While the reduced order model
structure has been repeatedly demonstrated to accurately represent the reactions taking place during
the industrial fermentation of beer, model parameters are subject to variations due to biological
system-specific factors (yeast strain, mutations, species aggregation). In our study the original model
parameters have been used (de Andrés-Toro, 1998); these have been reported as computed from
industrial data obtained from a fermentation campaign performed at a full scale industrial plant (Cruz
Campo Brewery, Madrid, Spain).

2.3 Literature review


Numerous authors have used the de Andrés-Toro (1998) beer fermentation model for optimal control
studies. Several have been stochastic approaches, including genetic algorithms (Carrillo-Ureta et al.,
2001), ant colony system (Xiao et al., 2003) and simulated annealing (Rodman and Gerogiorgis,
2016a). Additionally, Bosse and Griewank (2014) have used the kinetic model to generate optimal
control profiles using a sweeping dynamic optimisation methodology. The process involves guessing
a control path and using this to integrate the states forward in time. This allows the costates to be

7
integrated backwards through the process time span: a new control profile is thus deduced by
maximising the Hamiltonian for all t ∈ [𝑡0 , 𝑡𝑓 ], and the process is repeated until path convergence is
attained. The authors were able to compute a more preferable temperature profile using the same
objective, compared to a prior stochastic approach (de Andrés-Toro et al., 1997).

2.4 Process targets: objective function

When considering what it is desirable to improve in a fermentation process there are two obvious
contenders: reduced duration and heightened alcohol content (even if this requires later dilution, it is
still desirable to increase yield). In addition to batch time minimisation and alcohol production
maximisation, all prior authors have elected to include terms for minimisation of both by-products
within their respective optimisation objective functions. However, as is known that within certain beer
products the concentrations of both ethyl acetate and diactyl compounds shall be indistinguishable
below certain levels, efforts towards further reduction and concentration minimisation are redundant.
As such it is deemed more appropriate to consider an objective function only seeking to minimise
production time and maximise sugar conversion to ethanol (with variable relative weights) while
treating the final concentrations of both ethyl acetate and diacetyl compounds as strict constraints to
avoid unnecessary efforts towards further by product reduction.

Thus the following objective shall be used in this study:

̃
̃ ] − 𝑊𝑡 ∙ (1 )
min 𝜑(𝒙, 𝑡𝑓 ) = −𝑊𝐸 ∙ [𝐸𝑡𝑂𝐻 (18)
𝑇(𝑡),𝑡𝑓 𝑡
𝑠. 𝑡. [𝐸𝐴]𝑡=𝑡𝑓 ≤ [𝐸𝐴]𝑚𝑎𝑥 (19)
[𝐷𝑌]𝑡=𝑡𝑓 ≤ [𝐷𝑌]𝑚𝑎𝑥 (20)

Here 𝑊𝐸 and 𝑊𝑡 are the respective weights of the two components in the objective function: while a
large range of weight values have been investigated, for conciseness only the case 𝑊𝐸 = 𝑊𝑡 = 50%
̃
1
shall be presented in this paper. ( 𝑡 ) is the inverse batch time normalised by division with the

maximum value attained from exhaustive simulation (Rodman and Gerogiorgis, 2016a) and
̃ is the ethanol concentration normalised in the same way. In doing so the normalised ethanol
[𝐸𝑡𝑂𝐻]
̃ ranges from 0.68 when [𝐸𝑡𝑂𝐻] = 42 g L-1 to 1 when [𝐸𝑡𝑂𝐻] = 61.3 g L-1,
concentration, [𝐸𝑡𝑂𝐻]
̃
1
similarly the normalised inverse batch time, ( 𝑡 ), ranges from 0.62 to 1 when t is 99 hrs and 160 hrs

respectively.

Given the strong dependence of yeast health on system temperature it is necessary to include an
additional constraint such that the control profile (temperature manipulation schedule) remains within
acceptable levels. Eq. 21 ensures that the lower temperature limit excludes scenarios in which the

8
system lacks enough energy to promote cell growth while the upper limit ensures bacteria which are
present above this temperature cannot thrive, while also preventing the temperature from reaching a
level at which undesirably high by-product concentrations are known to be produced.

𝑇(𝑡) ∈ [9 °C, 16 °C] for all t ∈ [𝑡0 , 𝑡𝑓 ] (21)

2.5 Temperature profile: Performance insight


A prior exhaustive simulation campaign has been performed, considering a finite set (175,000) of
piecewise linear temperature manipulation profiles adhering to a low level of discretising with equally
spaced time intervals, which also adhere to realistic operability heuristics for control profile
formulation in order to reduce the feasible set considered (Rodman & Gerogiorgis, 2016a). To gauge
what quantifies as effective fermentation performance the solution set can be examined. Fig. 5
presents the performance of each candidate profile used for simulation, considering four metrics:
batch time (x-axis) as well as ethanol, diacetyl and ethyl acetate concentrations on the y-axis such that
each point corresponds to the performance of a single candidate profile. Here yellow markers indicate
solutions attaining an ethanol concentration above the prescribed minimum for each column, while
orange markers represent those which also fall below the base case limit on ethyl acetate
concentration, defined as [EA]max = 2.0 ppm. It is shown from the first column of Fig. 5 that profiles
producing ethanol concentrations above 60 g L-1 are universally unable to fulfil the tolerable base case
ethyl acetate threshold. Decreasing the acceptable ethanol concentration to 59.5 g L-1 in the middle
column shows that there is now a subset of profiles which produce by-product concentrations below
both [EA]max and [DY]max (0.1 ppm) for the base case, however it is found that these correspond to
unfavourably long batch times. Further decreasing the ethanol limit to 59 g L-1 shows in the right most
column that there is now a subset of solutions which correspond to short batch times (tf < 120 hr)
which attain this ethanol concentration while maintaining by-products below base case thresholds.

3. DYNAMIC OPTIMISATION OF BEER FERMENTATION


A wide range of optimisation methodologies exist for solving optimal control trajectory problems.
These include variation methods and finite approximation methods. In the former exploiting
Pontryagin’s maximum principle allows the resulting two point boundary value problem to be solved,
while the later uses predefined functional forms to represent the control profile (Almeida and Secchi,
2011). Finite formulations may be tackled with simultaneous, sequential or multi shooting strategies
which are extensively reviewed in the literature (Biegler et al., 2002). The sequential strategy involves
discretisation of the control profile with the ODE system (process model), requiring regular re-
integration during the algorithm to compute corresponding state trajectories, an approach effective for
problems with few decision variables and constraints (Osorio et al., 2005) which has been widely

9
applied to engineering problems (Farhat et al., 1990; Mujtaba and Macchietto, 1993; Sørensen et al.,
1996). In contrast, simultaneous strategies require the ODE system to also be discretised on the time
horizon to produce a large scale NLP problem requiring no futher integration of the DEA system,
generally using orthogonal collocation techniques. The later offers numerous benefits, being faster to
solve and able to handle problems with a greater number of decision variables and constraints
(Cervantes et al., 1998; 2000).

In this paper we have applied a simultaneous dynamic optimization strategy using collocation on
finite elements, which is an established methodology widely used in numerous chemical process
optimization studies: the key contributions advancing the state of the art is that we are pursuing
dynamic optimization for a combination of two competing objectives, and even more so under explicit
constraint level variation. The orthogonal collocation on finite elements is a trusted and robust
approach, which has been demonstrated to be applicable for constrained problems such as industrial
fermentation optimization, unlike variational methods which are not efficient for solving constrained
problems. Discretizing both the state and control variables to form large-scale NLPs allows rapid
determination of solution profiles with fewer finite elements than when using sequential methods that
apply standard ODE solvers and are unable to properly handle problem instabilities. The simultaneous
strategy has further advantages for the treatment of path constraints, useful if we were to extend the
problem, for example, to prohibit by-product constraint violation at any time in place of at terminal
time only. Limitations of the strategy employed are that efficient large scale NLP solvers are required,
while global optimization is cumbersome (indeed, the optimal solution profiles computed here are
indeed strongly dependent on the initialization considered, but they also clearly illustrate the
multitude of attainable tradeoffs between ethanol maximization and batch duration minimization).

3.1 DynOpt for fermentation optimisation


A direct method for dynamic optimisation (simultaneous strategy) has been performed in this study.
Orthogonal polynomials on finite elements are used to approximate the control and state trajectories
allowing the continuous problem described by Eqs. (6-21) to be converted to NLP form.
Implementation has been performed using the DynOpt package for MATLAB (Cizniar et al., 2006).
The DAE system is converted to a system of algebraic equations, where decision variables of the
derived NLP problem are the coefficients of the linear combinations of these AEs. Precision is known
to vary with collocation point locations and step sizes used (Logsdon and Biegler, 1989; Tanartkit and
Biegler, 1995).

Considering the general problem (Eqs. 1-5) with N elements (i = 1, …, N), each of which with K
collocation points (j = 1, …, K). The differential profiles (Eq. 2) can be approximated by:

10
𝑡−𝑡𝑖−1 𝑑𝑥
𝑥𝑖 = 𝑥𝑖−1 + ∆𝑡𝑖 ∑𝐾
𝑗=1 𝛺𝑗 ( ) (22)
∆𝑡𝑖 𝑑𝑡𝑖,𝑗

Where ∆𝑡𝑖 is the length of element i and 𝑑𝑥/𝑑𝑡𝑖,𝑗 is the derivative of the state variable in element i at
the jth collocation point. 𝛺𝑗 is a Kth order polynomial satisfying:

𝛺𝑗 (0) = 0, 𝛺′𝑗 (𝜌𝑗 ) = 𝛿𝑗 for 𝑗 = 1, … , 𝐾 (23)

Continuity of the state trajectories is ensured with:

𝑑𝑥
𝑥(𝑡) = 𝑥𝑖−1 + ∆𝑡𝑖 ∑𝐾
𝑗=1 𝛺𝑗 (1) (24)
𝑑𝑡𝑖,𝑗

while the control profile is approximated by:

𝑡−𝑡𝑖−1
𝑢(𝑡) = ∑𝐾
𝑞=1 𝜓𝑗 ( ) 𝑢𝑖,𝑗 (25)
∆𝑡𝑖

Where 𝜓𝑗 is a Lagrange polynomial of degree K that satisfies 𝜓𝑗 (𝜌𝑗 ) = 𝛿𝑗 for 𝑗 = 1, … , 𝐾.


It is shown in Fig. 6 how control variables may have discontinuities at element boundaries, while Eq.
24 produces continuity in states at these same boundaries.
Applying Eqs. 22-25 to the fermentation problem described by Eqs. 6 – 21 the resulting NLP problem
is as follows, where x is a vector containing the 7 model states (Eq. 6-12) and 𝑇𝑖,𝑗 is the fermenter
temperature in element i at collocation point j:
̃
min ̃ 𝑁 − 𝑊𝑡 ∙ ( 1 )
𝜑(𝒙, 𝑡𝑓 ) = −𝑊𝐸 ∙ [𝐸𝑡𝑂𝐻] (26)
𝒙𝑖,𝑗 , 𝑇𝑖,𝑗 ,∆𝑡𝑖 𝑡𝑓

𝑑𝒙
s.t = 𝑓(𝒙𝑖,𝑗 , 𝑇𝑖,𝑗 ) (27)
𝑑𝑡𝑖,𝑗

𝑑𝒙
𝒙𝑖,𝑗 = 𝑥𝑖−1 + ∆𝑡𝑖 ∑𝐾
𝑗=1 𝛺𝑗 ′(𝜌𝑗 ) (28)
𝑑𝑡𝑖,𝑗

ℎ(𝒙𝑖,𝑗 , 𝑇𝑖,𝑗 ) = 0, i = 1, …, N, j = 1, …, K (29)

𝑑𝒙
𝒙𝑖,𝑗 = 𝒙𝑖−1 + ∆𝑡𝑖 ∑𝐾
𝑗=1 𝛺𝑗 (1) , i = 1, …, N (30)
𝑑𝑡𝑖,𝑗

𝑔𝑓 (𝒙𝑁 ) ≤ 0 → [𝐸𝐴]𝑁 − [𝐸𝐴]𝑚𝑎𝑥 ≤ 0, [𝐷𝑌]𝑁 − [𝐷𝑌]𝑚𝑎𝑥 ≤ 0 (31)

9 ≤ 𝑇𝑖,𝑗 ≤ 16, 0 ≤ 𝒙𝑖,𝑗 , i = 1, …, N, j = 1, …, K (32)

𝑡𝑓 = ∑𝑁
𝑖=1 ∆𝑡𝑖 (33)

This large scale NLP problem produced from the DynOpt code has been solved with the fmincon
MATLAB function, using the interior point algorithm from the optimisation toolbox, a detailed
derivation of which is given by Waltz (2006). Three collocation points have been used for state
trajectories, with one collocation point being used for control profiles, resulting in the computation of
temperature profiles which are piecewise-constant. The NLP solver has been executed in each case for

11
a fixed number of iterations, rather than setting solution tolerances as stopping criteria: the maximum
function evaluations for each discretisation level is given in Table 2.

3.2 Initialization
Due to the high number of local extrema that exist when discretizing a control vector problem to NLP
form, the initializing profile has considerable bearing on the resultant output profile, which cannot be
guaranteed as globally optimal. An investigative campaign was performed using five isothermal
profiles to initialize the solver (T = 11, 12, 13, 14, 15 °C). As these isothermal profiles do not show
particularly suitable performance for industrial beer fermentation, the result was that the outputs in
fact did not represent profiles for great performance either, due to confinement to local solutions in
the vicinity of the isothermal input. To overcome this limitation, it is desirable to input a profile
known to have good performance, such that the algorithm can act to improve on this. We have
selected a range of profiles from a prior exhaustive simulation campaign (Rodman & Gerogiorgis
2016a) performed with a low discretization level, N = 6.
These corresponding profiles which have been used for initializing the DynOpt code, are shown in Fig.
7, where their position from Fig. 5 is highlighted with stars of corresponding colour in the plot which
is cropped around the desirable region. To visualise how these profiles perform, Fig. 8 depicts the
position of these points on the performance plots for the entire solution set. The top plot (ethanol vs.
batch time) shows that the four profiles taken forward from exhaustive simulation for initialising the
simultaneous optimisation procedure all fall towards the more desirable portion of the plot.
The vast number of points which correspond to lesser batch time and greater ethanol concentration
(top left corner of plot) however suggest that there is significant scope to improve upon these profiles.
The lower two plots (by-products vs. batch time) show that profiles A, C & D all universally fulfil
both base limits of the by-product species, while prolife B in fact does violate the diacetyl limit. This
is of interest to observe how a constraint violation in the initializing solution affects the performance
of the algorithm in producing optimal T(t) profile outputs.

As these initializing profiles are piecewise linear and DynOpt is computing piecewise constant
temperature profiles it is necessary to approximate the profiles in Fig. 7 to a piecewise constant form,
which will differ for each discretisation level solved. This approximation is performed by averaging
the temperature over N steps of equal duration: this transformation is shown in Fig. 9 for profile D. It
is demonstrated that N increases the profile tends to the original piecewise linear form.

4. RESULTS & DISCUSSION


To thoroughly investigate effect which by-product constraint thresholds (Eqs. 19-20) have the
attainable fermentation performance, and to access the methodology performance, a large campaign
of cases have been solved. Five realistic thresholds for ethyl acetate and diacetyl have been selected

12
producing 25 constraint permutations for which the system will be solved. Each is performed for eight
different discretisation levels and initialized with each of the four input profiles (Fig. 7) in turn,
meaning a solution set of 6400 control profiles has been produced, summarized in Table 2. A sample
of solution profiles are presented here for conciseness.

Fig. 10 shows the temperature profiles computed for the base case by-product concentration
thresholds, for each initializing profile and a selection of the discretization levels used. Figures 11-13
shows the corresponding profiles for different constraint threshold levels as indicated by the figure
captions: low [EA]max & high [DY]max; high [EA]max & low [DY]max; low [EA]max & low [DY]max
respectively, such that the Figs. 10 – 13 represent 4 of the 25 total permutations computed. Within the
final column of each figure (N = 42) the best and worst performing profiles for each aspect of the bi-
criteria objective are highlighted with the corresponding performance metric value within the plot
panel. Remarkably several profiles produced have a strong resemblance to the profile form obtained
by Bosse and Griewank (2014), showing characteristic dual peaks with a moderate dip in temperature
between them. The authors did use the same fermentation model, however described a quite different
objective, considering [EA] and [DY] as minimisation criteria rather than constraints as done here, in
addition to not considering batch time as a minimisation target.

4.1 Effect of increasing time domain discretisation

Inspecting horizontal rows from Figs. 10 – 13 in isolation allows the influence of the discretisation
level to be observed. It is apparent that regardless of the discretisation level the solution form is
similar: in general as N increases the profile becomes refined, following a similar trajectory in a
smoother manner. There are however several instances where considerable deviations occur. Firstly
from Figure. 10 the upper row (input profile A) shows that depending on the discretisation level, the
solution profile has a drastically different initial temperature, T(t=0), ranging from 11 to 16 °C. A
similar observation can be made from the third row of the same figure (input profile C). The case
presented in Figure 11 also shows considerable differences in the profile as N increases. At lower
levels of N the solution produced is very flat, while once a greater degree of control is allowed a much
more variable solution is obtained, which corresponds to a significantly improved objective. A further
example of the solution being sensitive the discretisation level is shown in Figure 12 row 4 (input
profile D). The profiles initial temperature, T(t=0), is considerably lower at high values of N
compared to that at lower discretisation levels. These differences indicate the discretisation level can
have significant bearing on the specific profile being produced, however in most cases the overall

13
solution form does not differ drastically. There are some cases in which lower discretization levels
yield shorter batch times compared to those obtained for higher discretization levels (e.g. Figure 13,
cases A and C); these occurences are affected by both the initialisation T(t) profile, but also by the
fact that gradual temporal grid refinement favors further [EtOH] maximization at the expense of batch
duration (tf) minimization.

It is demonstrated that as the number of piecewise constant sections in the profile exceeds 40 the
solution is smoothed and tends towards a continuous form, which ultimately removes the
implementation issues of using piecewise constant profiles (instantaneous temperature adjustments). It
is shown that as N is increased the attainable value for the objective function (Eq. 18) increases, as
expected due to the greater level of control possible with a higher number of manipulatible sections in
the temperature profile.

4.2 Effect of initialising temperature profile

To observe the influence of the initializing profile it is necessary to inspect columns from Figs. 10 –
13 in isolation. An immediate observation is that solutions do not converge to the same solution,
meaning that globally optimal solutions are not being produced, rather the input initialising profile has
significant impact on the profile output for any set of conditions when using this methodology.
However, there exists a large number of similarities between the solution profiles’ appearance, with
significant features present across all solutions even when differences in the duration or magnitude of
these features exists. Comparing the performance of solutions which differ only in the initialising
profile (each figure column) show very similar values, with the objective value (Eq. 18) only differing
by more than 1% in very few cases. Within Figure 11. The profile produced at N = 42 when initialized
with profile D is considerably different from those from the alternative input profiles. Its secondary
peak is not present in the other profiles which significantly affects performance. A lower ethanol
concentration is produced than all three other solutions, however batch time is drastically less such
that case D has the most desirable objective. This is likely a result of initializing profile D featuring
this late peak, which does correspond to the best performing of the 4 input profiles considered.
Furthermore, within Figure 13 the final columns (N = 42) shows two differing solution forms: input
profiles B and D produce an output profile with a gradual temperature reduction followed by a later
peak, while inputs A and C produce an alternative form with a gradual temperature increase towards
this peak. The former solution form corresponds to a greater ethanol production, while the later
permits a much shorter batch duration, however once again the overall objective is not drastically
different. This emphasises the influence which the input profile has on the resultant solution obtained
by the algorithm, given that globally optimality is not being achieved.

4.3 Effect of by-product constraint thresholds

14
To discover the influence which the concentration limits of by-product species in the beer product
have the attainable fermentation efficiency it is necessary visualise the performance of all 25
threshold permutations simultaneously. Figure 14 presents this for the discretization level N = 30 with
the four plots corresponding to the results using the four different initializing profiles. Each point
corresponds the performance an output profile; the corresponding by-product concentration limits are
represented on the x-y plane with the z axis showing the batch time, tf and the marker colour
corresponding to the ethanol product concentration, [EtOH]tf.

It is observed that for all input profiles, the resultant performance points are very close, reiterating that
while the solutions do differ their performance is highly comparable across the initialling profiles
used. The results show a very coherent pattern indicating the manner in which fermentation
performance is influenced. It is shown that batch time universally increases as the acceptable
threshold on diacetyl, [DY]max, is reduced. Batch time does indeed also increase as [EA]max is reduced,
however the relationship is far less significant, with the dependency on the diacetyl threshold much
stronger. The marker colours show how it is exclusively the ethyl acetate threshold, [EA] max, which
influences the final ethanol yield. In all cases when [EA]max, = 0.5 ppm the product ethanol, [EtOH]tf,
is very low (under 56 g L-1), which increases steadily towards 61 g L-1 as this permitted [EA]max,
threshold is relaxed towards 3 ppm. These results reveal as to how the two components of the bi-
criteria objective are dictated by the two inequality constraints on the by-product concentrations:
 [DY]max has very strong influence on batch time.
 [EA]max is shown to dictate the attainable ethanol concentration.

Fig 15. represents the projection of the four plots from Fig 14. on the by-product concentration plane.
Here the marker size is scaled relative to inverse batch times: smaller makers show the worst
performing solutions (longest batch times) while larger markers show the best performance (shortest
batch times). It is shown that a small selection of solutions do not fall directly on the intersect of the
two by-product limits imposed for the particular case being solved. There are two factors responsible
for this; firstly there are several cases where the constraints are comfortably fulfilled, ie the solution
produced for the case [EA] < 1.0 ppm & [DY] < 0.25 ppm in fact has a much lower [DY] of only 0.22
ppm. This could suggest sub optimality in the solution, perhaps with a shorter batch time possible if
the concentration of diacetyl were to increase more. Secondly, all performance results presented in
this work have been computed after reintegration of the system using the solution control profile.
Slight deviations exist between the performance of the profile during the NLP algorithm and later
integration of the solution, depending on the accuracy of the piecewise polynomial representation of
the continuous state trajectories in the NLP formulation. Deviations have are shown to be non-
significant as the algorithm used captures the state trajectories effectively.

4.4 Performance of key output profiles

15
Of the entire solution set computed it is of interest to inspect the profiles which correspond to the
extrema in terms of performance. As such 5 cases are presented from the N = 30 solution set:

I. The best performing profile from the base case ([EA] max = 2.0 ppm, [DY]max = 0.10 ppm)
II. The profile corresponding to the longest batch time, tf, from all cases
III. The profile corresponding to greatest ethanol concentration, [EtOH]tf, from all cases
IV. The profile corresponding to the shortest batch time, tf, from all cases
V. The profile corresponding lowest ethanol concentration, [EtOH]tf, from all cases

To allow visualization of how these compare among the rest of the solutions, they are highlighted in
the collated performance plot depicted in Figure 16. The corresponding profiles are shown in Figure.
17, as well as the state trajectories of all species considered in the dynamic model. The performance
of these profiles is summarised in Table 3.

It is demonstrated that the solution which corresponds to the shortest batch times aligns with the
greatest value of [DY]max, and conversely the longest batch corresponds to the maximum value of
[DY]max, further confirming the strong correlation between the two parameters. It is shown that the
greatest product ethanol concentration occurs when both constraints are fully relaxed, with the
lowest ethanol concentration corresponding to the tightest by-product limits, indicating that both
species thresholds in fact influence the obtainable ethanol concentration. It is noteworthy that the
solutions corresponding to desirable results (I, III, IV) are produced using initializing profile D; this
input profile is the scenario which maximises Eq. 18 from the entire exhaustive simulation set
(Rodman & Gerogiorgis, 2016a). Similarly the two undesirable solutions representing minimum
ethanol and longest batch times (II, V) have been computed from profile B which is the least
preferable of the 4 input profiles considered. This demonstrates that a better performing input profile
enables the computation of the most preferable output profiles from the DynOpt algorithm, given the
solution sensitivity to the input profile.

5. CONCLUSIONS
A multi-objective dynamic optimisation study has been carried out, investigating the potential for
process improvement of industrial scale batch beer fermentation via modifications to the fermentor
temperature profile throughout the duration of the process stage. A simultaneous method for direct
dynamic optimisation of the temperature profile has been performed for a spectrum of threshold
values on by-product concentrations to investigate the effect which these have on the obtainable
process performance.

The scope and demand for beer fermentation optimisation has been motivated and justified on the
basis of a detailed survey of production statistics, and optimal fermentor temperature manipulations

16
(dynamic profiles) towards minimising batch duration and maximising product yield (final ethanol
concentration) have been computed via a simultaneous strategy. A key contribution of the present
study is the explicit consideration of by-product constraint level variability, as the extensive
illustration and comparative evaluation of optimal T(t) profiles for constraint variation (ethyl acetate,
[EA] and diacetyl, [DY] thresholds) clearly reveal the pivotal influence that these limits have on
processing targets. The maximum allowable ethyl acetate concentration significantly affects the
ethanol yield: a relaxation from 0.5 ppm to 3 ppm increases ethanol from 55.5 g L-1 to 65.5 g L-1.The
maximum allowable diacetyl concentration in the final product also has a very strong influence, but
mainly on batch duration: a relaxation from 0.05 ppm to 0.25 ppm can reduce batch time by up to
33%.

The sequential dynamic optimisation procedure has been performed using the DynOpt package for
MATLAB. Discretising the state trajectories in addition to the control vector using orthogonal
collocations permits a large scale NLP problem to be solved, here for piecewise constant temperature
profiles. This is performed for 25 different pairs of constraint thresholds on ethyl acetate and diacetyl
in the beer product to compare the performance of the optimal profiles produced in each case. A range
of discretization levels (N) and initializing profiles are used for each scenario producing a total of
6400 unique cases which have been solved.

The implemented algorithm does not address global optimisation, exactly because of the clear and
prominent tradeoffs between the two objectives (ethanol maximisation and batch duration
minimisation), but also due to the sensitivity of optimal T(t) solution profiles which depend on the
initialization profile T0(t) considered. It is shown that the best performing input profiles produce the
best performing outputs, however the difference in performance is not significant providing an
acceptable profile is used for initialisation. It is shown that as N is increased the attainable value for
the objective improves, as expected due to the greater level of control possible with a higher number
of manipulatible sections in the temperature profile. It is demonstrated that as the number of
piecewise constant sections in the profile exceeds 40 the solution is smoothed and tends towards a
continuous form, which ultimately removes the implacability of using piecewise constant profiles
(instantaneous temperature adjustments).

The investigation into the influence of by-product threshold limits on obtainable fermentation
performance has revealed new insight into how each by-product uniquely affects process performance.
It is found that the permitted diacetyl concentration in the product has very strong influence on batch
time, with lower limits requiring considerably longer batches. Ethyl acetate is shown to dictate the
attainable ethanol concentration, such that low limits prohibit a reasonable alcohol content in the
product.

17
ACKNOWLEDGEMENTS
The authors gratefully acknowledge the financial support of the Eric Birse Charitable Trust for a Birse
Doctoral Fellowship awarded to Mr A.D. Rodman, and that of the Engineering and Physical Sciences
Research Council (EPSRC) via funding from an Impact Acceleration Account (IAA) administered by
Edinburgh Research & Innovation (ERI). Moreover, Dr D.I. Gerogiorgis gratefully acknowledges a
Royal Academy of Engineering (RAEng) Industrial Fellowship which he has been awarded (2017).
Both authors express thanks to Mrs Hilary Jones, Mr Simon P. Roberts and Mr Udo Zimmermann
(WEST Beer) for consistent encouragement and inspiring discussions throughout this research project.

18
REFERENCES
1. Almeida Nt. E. and Secchi, A. R., 2011. Dynamic optimization of a FCC converter unit: numerical
analysis. Brazilian Journal of Chemical Engineering, 28(1): 117-136.
2. Beer Statistics (2015) – Available online at: www.brewersofeurope.org/uploads/mycms-
files/documents/publications/2015/statistics_2015_v3.pdf.
3. Biegler, L. T., 2006. An overview of simultaneous strategies for dynamic optimisation. Chemical
Engineering and Processing, 46: 1043-1053.
4. Biegler, L. T., 2010. Nonlinear programming: concepts, algorithms, and applications to chemical
processes. SIAM publishing. Print.
5. Biegler, L. T., Campbell, S. L. and Mehrmann, V., 2012. Control and optimization with differential-
algebraic constraints. SIAM publishing. Print.
6. Biegler, L. T., Cervantes, A. M. and Wächter, A., 2002. Advances in simultaneous strategies for
dynamic process optimization. Optimization, Chemical Engineering Science, 57: 575-593.
7. Bonvin, D., 1998. Optimal operation of batch reactors—a personal view. Journal of Process
Control, 8(5): 355-368.
8. Bosse, T. and Griewank, A., 2014. Optimal control of beer fermentation processes with Lipschitz‐
constraint on the control. Journal of the Institute of Brewing, 120(4): 444-458.
9. Carrillo-Ureta, G., Roberts, P. and Becerra, V., 2001. Genetic algorithms for optimal control of beer
fermentation. Proceedings of the IEEE International Symposium on Intelligent Control, 391-396.
10. Cervantes, A. and Biegler, L. T., 1998. Large-scale DAE optimization using a simultaneous NLP
formulation. AIChE Journal, 44(5): 1038-1050.
11. Cervantes, A., Wachter, A., Tutuncu, R. H. and Biegler, L. T., 2000. A reduced space interior point
strategy for optimization of differential algebraic systems. Computers and Chemical Engineering, 24:
39-51.
12. Cizniar, M., Fikar M, and Latifi, M. A., 2006. MATLAB Dynamic Optimisation Code DYNOPT,
User's guide, Technical Report, KIRP FCHPT STU, Bratislava.
13. de Andrés-Toro, B., Giron-Sierra, J., Lopez-Orozco, J. and Fernandez-Conde, C., 1997. Application
of genetic algorithms and simulations for the optimization of batch fermentation control. 1997 IEEE
International Conference on Systems, Man, and Cybernetics. Computational Cybernetics and
Simulation, 392-397.
14. de Andrés-Toro, B., Giron-Sierra, J., Lopez-Orozco, J., and Fernandez-Conde, C., 1998. A kinetic
model for beer production under industrial operational conditions. Mathematics and Computers in
Simulation, 48(1): 65-74.
15. ence on Decision and Control, Vol. 1, pp. 828–829.
16. Farhat, S., Czernicki, M., Pibouleau, L. and Domenech, S., 1990. Optimization of multiple-fraction
batch distillation by nonlinear programming. AIChE Journal. 36: 1349-1360.
17. Gee, D. A. and Ramirez, W. F., 1988. Optimal temperature control for batch beer fermentation.
Biotechnology and Bioengineering, 31: 224-234.

19
18. Hanke, S., Ditz, V., Herrmann, M., Back, W., Becker, T and Krottenthaler, M., 2010. Influence of
ethyl acetate, isoamyl acetate and linalool on off-flavour perception in beer. Brewing Science, 63(7):
94-99.
19. Izquierdo-Ferrero, J. M., Fernández-Romero, J. M., Luque de Castro, M. D., 1997. On-line flow
injection–pervaporation of beer samples for the determination of diacetyl. The Analyst, 122(2): 119-
122.
20. Logsdon, J. S. and Biegler, L. T., 1989. Accurate solution of differential-algebraic optimization problems.
Industrial and Engineering Chemistry Research, 28(11): 1628-1639.
21. Mujtaba, I. M. and Macchietto, S., 1993. Optimal operation of multicomponent batch distillation-
multiperiod formulation and solution. Computers and Chemical Engineering. 17(12): 1191-1207.
22. Osorio, D., Pérez-Correa, J. R., Biegler, L. T. and Agosin, E., 2005. Wine distillates: practical
operating recipe formulation for stills. Journal of agricultural and food chemistry, 53(16): 6326-6331.
23. Rodman, A. D. and Gerogiorgis. D. I., 2016a. Multi-objective process optimisation of beer
fermentation via dynamic simulation. Food and Bioproducts Processing, 100: 255-274.
24. Rodman, A. D. and Gerogiorgis, D. I., 2016b. Dynamic simulation and visualisation analysis of
fermentation: effect of conditions on beer quality, IFAC-PapersOnline (DYCOPS 2016), 49(7): 615-620.
25. Rodman, A. D. and Gerogiorgis, D. I., 2016c. Multi-objective optimisation of flavour and processing
time in beer fermentation via dynamic simulation, Computer-Aided Chemical Engineering, 38: 1033-
1038.
26. Sørensen, E., Macchietto, S., Stuart, G. and Skogestad, S., 1996. Optimal control and on-line
operation of reactive batch distillation. Computers and Chemical Engineering, 20(12): 1491-1498.
27. Tanartkit, P. and Biegler, L. T., 1995. Stable decomposition for dynamic optimization. Industrial and
Engineering Chemistry Research, 34(4): 1253-1266.
28. Trelea, I. C., Titica, M., Landaud, S., Latrille, E., Corrieu, G. and Cheruy, A., 2001. Predictive
modelling of brewing fermentation: from knowledge-based to black-box models. Mathematics and
Computers in Simulation, 56(4): 405-424.
29. Vanderhaegen, B., Neven, H., Verachtert, H. and Derdelinckx, G., 2006. The chemistry of beer aging
– a critical review. Food Chemistry, 95(3): 357-381.
30. Waltz, R. A., Morales, J. L., Nocedal, J. and Orban, D., 2006. An interior algorithm for nonlinear
optimization that combines line search and trust region steps, Mathematical Programming, 107(3):
391–408.
31. Xiao, J., Zhou, Z., Zhang, G., 2003. Ant colony system algorithm for the optimization of beer
fermentation control. Journal of Zhejiang University SCIENCE, 5(12): 1597-1603.

20
Figure 1. Alcohol consumption per capita, for UK adults (15+) (Beer Statistics, 2015).

Figure 2. Number of UK breweries in operation by year (Beer Statistics, 2015).

Figure 3. Alcohol consumption by country (Beer Statistics, 2015).

21
Figure 4. Kinetic model for beer fermentation under industrial conditions (de Andres-Toro, 1998).

[EtOH]min = 60 g L-1 [EtOH]min = 59.5 g L-1 [EtOH]min = 59 g L-1

Figure 5. Corresponding profile performance for lowering acceptable [EtOH]tf levels. Yellow staining of high
[EtOH]tf implies unacceptably high [EA]tf. Orange staining of acceptable [EA]tf indicates [DY]tf satisfaction.

22
Figure 6. Collocation method for state and control profiles (based on Biegler, 2006).

Figure 7. Promising profiles from exhaustive simulation to be used for initilaization of DynOpt.

23
Figure 8. Performance of four highlighted profiles with respect to by-product concentration thresholds.

Figure 9. Piecewise constant approximations of profile D (Fig. 7) for varying discretisation levels.

24
Figure 10. Computed profiles for base case: [EA] max = 2.0 ppm, [DY]max = 0.10 ppm.

Figure 11. Computed profiles for low [EA]max = 0.5 ppm, high [DY] max = 0.25 ppm.

25
Figure 12. Computed profiles for high [EA] max = 3.0 ppm, low [DY] max = 0.05 ppm.

Figure 13. Computed profiles for low [EA]max = 0.5 ppm, low [DY]max = 0.05 ppm.

26
Figure 14. Performance of output profiles (N = 30) for all constraint permutations and initialising profiles.

Figure 15. Performance plot projection on by-product threshold plane (N = 30): marker size scaled with t f-1.

27
Figure 16. Collated profile performance (N = 30) for all constraint permutations and initialising profiles.

Figure 17. State trajectories corresponding to key temperature profiles computed.

28
Table 1. Model parameters for dynamic simulation of beer fermentation

Symbol Description 𝐕𝐚𝐥𝐮𝐞 Units


Diacetyl rates 1.27672∙10
𝜇𝐷𝑌 Diacetyl production rate
-7
(Carrillo-Ureta et al., 2001) g-1 h-1 L
Diacetyl consumption rate 1.13864∙10
𝜇𝐴𝐵
-3
g-1 h-1 L
[XA]+[XL]+[XD] Biomass inoculum (pitching rate) 4 g L-1
Initial simulation conditions
[𝑆]0 Sugar concentration 130 g L-1

Table 2. Summary of solution conditions, producing 800 cases

Number of values Range


[EA]max (ppm) 5 [0.5, 1.0, 1.5, 2.0, 3.0]
[DY]max (ppm) 5 [0.05, 0.10. 0.15, 0.20, 0.25]
Discretisation level, N 8 [6, 12, 18, 24, 30, 36, 42, 48]
Initializing profile 4 [A, B, C, D]
Max function evaluations for N [2000, 5000, 10000, 15000,
20000, 25000, 30000, 30000]

Table 3. Performance of extrema profiles computed.

Solution Initialising profile [EtOH]tf (g L-1) tf (hr) [EA]max (ppm) [DY]max (ppm)
I D 59.8 114.2 2.0 0.10
II B 59.6 142.9 3.0 0.05
III D 60.5 128.5 3.0 0.05
IV D 60.1 94.9 3.0 0.25
V B 55.4 120.3 0.5 0.05

29

You might also like