You are on page 1of 29

Kinetic and Process Modeling of UV-C Irradiation of Foods

Mehmet R Atilgana, Semanur Yildizb, Zehra Kayac, and Sevcan Unluturkc, a Ebiltem TTO, Ege University, Izmir, Turkey;
b
Department of Food Engineering, Sakarya University, Sakarya, Turkey; and c Department of Food Engineering, Izmir Institute of
Technology, Izmir, Turkey
© 2020 Elsevier Inc. All rights reserved.

Introduction 1
Factors Affecting the Performance of UV-C Processing of Foods 2
Microbial Inactivation by UV-C Light 3
Modeling of UV-C Inactivation Kinetics 3
Log-Linear Models 3
Weibull Model 4
Hom Model 5
Log Logistic Model 5
Gompertz Model 5
Biphasic Model 5
Baranyi Model 5
Kinetic Power Law Model 6
Modeling UV-C Induced Inactivation of Microorganisms on Food Surfaces 6
Modeling UV-C Induced Inactivation of Microorganisms in Liquid Foods 6
Combined UV-C Treatment With Other Technologies 11
Modeling Microbial Inactivation Kinetics of Combined UV-C and Heat Treatments 11
Mathematical Modeling of UV-C Processing of Foods Using Computational Fluid Dynamics (CFD) 14
Performing a CFD Analysis 14
Simulation of UV-C Intensity Distribution on Solid Food Surfaces 17
Simulation of UV-C Intensity Distribution in Liquid Foods 18
Concluding Remarks 25
References 26
Further Reading 29

Introduction

Food preservation is a process preventing the deterioration and spoilage of food products; it extends their shelf-life, while assuring
consumers a product free of pathogenic microorganisms. There are traditional and modern methods widely utilized to preserve
foods. The most common is the thermal method performed at high temperatures. However, heating damages the nutritional and
sensory quality of foods, for example, causing permanent loss of flavor and taste, breakdown of nutrients and resulting undesirable
browning reactions. Thus, it has a negative impact on the quality of foods (Garde-Cerdán et al., 2007; Walkling-Ribeiro et al., 2008).
Ultraviolet irradiation has been approved as a novel technology for the treatment of fruit juices (U.S. FDA, 2000). The US FDA
and USDA concluded that the usage of ultraviolet light at 253.7 nm (UV-C) for food processing is safe and its usage for reduction of
pathogens and deteriorative microorganisms is approved as an alternative treatment to thermal processing (Shah et al., 2016).
UV-C irradiation is known to be environmentally friendly and easy to install, safe and compatible with other technologies (Bint-
sis et al., 2000; Pereira and Vicente, 2010). However, the inactivation efficiency of the UV-C irradiation (or UV-C treatment) depends
on many factors including the resistance of microorganisms to UV-C light, the geometric configuration of UV-C reactor (geometry,
fluid depth, flow type, etc.), the operating conditions of the system (flow rate, exposure time, UV-C dose (fluence) etc.), and product
composition (physicochemical properties, e.g. color, suspended/soluble particles, absorbance, and initial microbial load etc.)
(Barbosa-Cánovas et al., 2005; Koutchma et al., 2004).
Mathematical models associated with inactivation kinetics may significantly reduce the number of challenge tests required to
determine, for example, inactivation efficiency of a UV-C reactor. Because, kinetic models allow the description and prediction
of microbial behavior under specific environmental conditions (Fakruddin et al., 2011). Additionally, they can help to quantify
the effects of intrinsic, extrinsic and/or processing factors on the resulting microbial proliferation in food products or food model
systems. Thus, the kinetic modeling is an important tool that can be used to optimize processing capacities and inactivation rates
(Ngadi et al., 2003).
Mathematical modeling can also be used to simulate UV-C light intensity distribution and flow field prediction in a UV-C
reactor. Simulation of a UV-C irradiation process based on the inactivation kinetics of target microorganism and the properties
of the liquid food can help to identify the location of the least treated liquid and the dead spot in the reactor (Ye and Koutchma,
2010). Thus, simulating the UV-C reactor using mathematical models is important in terms of creating a virtual prototype, reducing
costs by reducing design cycle times, and evaluating more design alternatives than is practically possible.

Reference Module in Food Sciences https://doi.org/10.1016/B978-0-08-100596-5.22972-7 1


2 Kinetic and Process Modeling of UV-C Irradiation of Foods

The important factors affecting UV-C treatment of foods, kinetic models used for prediction of the efficiency of microbial inac-
tivation, and mathematical models employed for estimation of UV-C dose distribution in the UV-C systems including integrated
numerical models to evaluate the performance of annular flow UV-C disinfection reactors are briefly described in this chapter.

Factors Affecting the Performance of UV-C Processing of Foods

The use of UV-C irradiation for disinfection of the surface of solid foods such as meat, fish and freshly cut fruits is limited by its
complex interaction with food structure and topography (Gayan et al., 2014a). The use of UV-C treatments for liquid foods is
more promising, especially for vegetable and fruit juices. However, the application of UV-C treatment to liquid foods is limited
due to the low penetration depth in liquid foods containing large amounts of UV-C absorbing compounds and suspended particles,
and the variation in residence time of the product in the UV-C reactor. Thus, the lethal efficacy of the UV-C treatment is reduced. It is
essential to design reactors that overcome low penetration depth and non-homogeneous dose distribution to ensure that all product
particles are exposed to UV-C light for a sufficient period of time and fluence (dose). For the design of such reactors, the effect of
product and process parameters influencing the delivery of adequate UV-C dose to liquid materials should be known and
understood.
These factors can be classified into three main categories as process and food product parameters as well as microbial character-
istics (Guerrero-Beltrán and Barbosa-Cánovas, 2004; Koutchma, 2014; Mañas and Pagán, 2005).
Several physicochemical (pH, total acidity, total soluble solid content, water activity, viscosity, density) and optical properties
of foods (absorption coefficient, turbidity, color) are significant environmental or extrinsic factors influencing the inactivation
efficiency of UV-C process. Guerrero-Beltrán and Barbosa-Cánovas (2005) stated that high total soluble solid content (Brix)
decreases the UV-C light penetration ability in the liquid foods. However, many studies showed that pH, total acidity and soluble
solid content of the food are not significantly affecting the efficiency of UV-C process (Hijnen et al., 2006; Koutchma et al., 2004;
Murakami et al., 2006; Ngadi et al., 2003). Density and viscosity of liquid foods are related to flow characteristics and determine
the flow behavior of fluid during UV-C processing. Viscous foods require more energy and special reactor design for an effective
inactivation (Koutchma, 2014).
Optical properties of foods are crucial factors for determination of the penetration ability and transmissivity of UV-C light. As the
UV-C light passes through the liquid, its intensity is reduced. Liquid foods can be classified as clear and opaque (cloudy) depend on
their absorption coefficient, turbidity and color properties. Colorless, clear and pure water has the highest UV-C penetration ability
(Guerrero-Beltrán and Barbosa-Cánovas, 2004). Liquid foods with a high absorption coefficient have very low UV-C penetration
depth and absorb UV-C light poorly comparing to clear liquids and water. It has been reported that 90% of the UV-C light can
pass through only the first 1 mm layer of the liquid foods (Guerrero-Beltrán and Barbosa-Cánovas, 2004).
The absorption coefficient, calculated from dividing absorbance with the path length, represents the magnitude of the UV-C
absorbance of the medium. It is considered a good indicator of the UV-C penetration depth in clear liquids with low turbidity.
The intensity of UV-C light absorbed by liquid food decrease by increasing absorption coefficient (Guerrero-Beltrán and
Barbosa-Cánovas, 2004). Thus, it is a major factor for determination of potential lethality of UV-C treatment (Keyser et al.,
2008; Sizer and Balasubramaniam, 1999).
Suspended or insoluble particles and the level of initial microbial load of liquid foods are other important factors affecting the
inactivation efficiency of UV-C irradiation treatment. Shielding of microorganisms by suspended particles leads to reduced UV-C
inactivation efficiency. Besides, high initial microbial load results in higher shielding/shadowing effect as well. Additionally, Gayan
et al. (2014a) pointed out that the growth phase, growing conditions, stress factors prior to UV-C treatment, and recovery conditions
after processing are among important extrinsic factors influencing UV-C microbial resistance. Suspended particles (minerals,
proteins, phenolic matter, starch etc.) absorb UV-C light themselves (causing tailing effect), scatter the light and inhibit UV-C light
to reach the microorganisms (Fenner and Komvuschara, 2005; Geeraerd et al., 2000; Keyser et al., 2008).
Color compounds can absorb, reflect, and scatter the incident light, reducing the number of photons available to inactivate
microorganisms (Koutchma et al., 2009). The presence of colored compounds and pulp particles causes poor UV-C light transmis-
sion resulting in low efficiency of UV-C disinfection (Char et al., 2010).
UV-C dose (UV-C fluence) and flow behavior are important intrinsic process parameters for the design of a UV-C system
(Koutchma, 2014). UV-C dose is generally defined as the amount of radiant energy delivered or received per unit area (Gayan
et al., 2014a). J$m2 or mJ$cm2 are commonly used units for UV-C dose (Koutchma et al., 2009) and calculated from the product
of UV-C intensity (W$m2) and exposure time (Retamar et al., 2019). UV-C intensity is virtually synonymous with the term UV-C
irradiance which is the total radiant energy incident on some surface from all directions. On the other hand, fluence rate is the UV-C
fluence across a surface, and the technically correct term that describes UV-C irradiance on non-flat surfaces. The UV-C intensity
greatly depends on the distance from the UV-C light source and the transmittance of the medium. Some authors have proposed
to express UV-C dose (D) as the total UV-C energy (radiant power) delivered to a volume flow, expressed in J$L1 or J$mL1 (Keyser
et al., 2008; Geveke, 2008; Müller et al., 2011).
The flow behavior of liquid foods, i.e., laminar or turbulent, is another factor necessary to be determined for an efficient UV-C
system design. The microorganisms’ trajectory through the irradiated zone in a reactor depends on the characteristics of flow and
liquids properties. An ideal UV-C reactor must provide perfect mixing (complete radial mixing (not longitudinal)) to ensure that all
parts of the liquid approach uniformly to the light source, so that all microorganisms receive the same dose. Turbulence with the
Kinetic and Process Modeling of UV-C Irradiation of Foods 3

high flow rates causes better mixing in the reactor and allows all parts of liquid to be exposed to the UV-C light more efficiently
(Barbosa-Cánovas et al., 2005; Guerrero-Beltrán and Barbosa-Cánovas, 2004). Although these ideal reactors do not exist, real units
should be designed to approximate these conditions. Because of the hydrodynamic behavior of a liquid within a reactor, the char-
acteristics of the UV-C light source and absorptive properties of the treated medium, it is important to estimate the UV-C intensity
distribution and the residence time distribution (RTD) to evaluate the performance of UV-C processing reactors. Computational
fluid dynamics (CFD) simulation software tools are available to compute the RTD and UV-C intensity distribution as well as parti-
cles and fluid velocities, and particle mixing in UV-C reactors (Ye and Koutchma, 2010).

Microbial Inactivation by UV-C Light

Pathogenic and spoilage microorganisms are responsible for foodborne diseases and quality losses that reduce the shelf life of
foods. Pathogenic bacteria such as E. coli, Listeria monocytogenes, Salmonella, B. cereus, Campylobacter jejuni, Candida botulinum,
Candida perfringens, A. hydrophila, Shigella, Staphylococcus aureus, Vibrio spp., Y. enterocolitica are major concerns related to foodborne
diseases and outbreaks (Koutchma et al., 2009). Microorganisms have different UV-C light sensitivity and require special lethal UV-
C dose. Many factors, such as cell wall structure, thickness and composition, the presence of UV-C absorbing proteins, and variation
in nucleic acid configuration, affect the susceptibility of the microbial cell to UV-C treatment (Barbosa-Cánovas et al., 2005;
Koutchma, 2014).

Microorganisms can be categorized in terms of their resistance to UV-C light as:


Gram negatives < Gram positives < yeasts < bacterial spores < molds < viruses (Barbosa-Cánovas et al., 2005). Johnson et al.
(2010) indicated that the smallest cells are the most resistant to UV-C light due to the amount of UV-C absorbed per cell.
However, yeasts and molds are also more resistant to UV-C light, because they have less pyrimidine bases and have different cell
membrane and thickness (Tran and Farid, 2004). Number of microorganisms and growth stage of microorganisms are also
important factors affecting the performance of UV-C process (Barbosa-Cánovas et al., 2005).

Modeling of UV-C Inactivation Kinetics

Inactivation kinetics can be described with the concept of “predictive modeling” based on several mathematical equations (Valdra-
midis et al., 2005). These equations may help to understand physiological behavior of microorganisms toward inactivation agents
and predict microbial inactivation in a wider range of conditions. They provide a better knowledge about the effect of a process on
the microorganisms and lead to more rational design of the process (Mañas and Pagán, 2005). Thus, the modeling of UV-C inac-
tivation kinetics is useful to estimate the efficiency of UV-C irradiation on the inactivation of microorganisms in foods; to predict
UV-C dose distribution in the system based on the absorptive and physicochemical properties of liquid foods; to establish appro-
priate UV-C processing conditions to achieve maximum level of microbial inactivation allowing the production of stable and safe
foods (Mañas and Pagán, 2005; Koutchma, 2014; Ye and Koutchma, 2010).
Generally, microorganisms in foods exposed to thermal and nonthermal treatments can exhibit different inactivation behaviors
described by a log-linear, linear with tailing, sigmoidal-like, linear with shoulder, biphasic types of curves etc (Barbosa-Cánovas
et al., 2005; Chen, 2007; Geeraerd et al., 2005; Peleg and Penchina, 2000). The inactivation kinetics of microorganisms depend
on several factors such as process parameters (time, temperature, dose etc.), types of microorganisms (bacteria, yeasts, molds
etc.) and food properties (pH, water activity, medium composition, absorbance, turbidity, color of liquids; surface morphology
of solids etc.) (Mañas and Pagán, 2005).
Linear model has been widely used for modeling of heat inactivation and novel technologies (Chen, 2007; Schaffner and Lab-
uza, 1997). However, survival curves are not always linear (Fig. 1). This is due to the occurrence of deviations such as shoulder and
tail (Mañas and Pagán, 2005). Shoulder effect has been mainly assigned to the occurrence of sub-lethal injury, multi-target inacti-
vation, cell clumping or activation phenomena for spores (Marugán et al., 2008). The tailing effect is considered to be due to the
heterogeneity of resistance in the population that is specific to bacterial cells or acquired during treatment (Bialka et al., 2008; Buzrul
and Alpas, 2007; Fernández et al., 2010; Huang, 2009; Marugán et al., 2008; Van Boekel, 2002). Tailing in UV-C inactivation curves
is also attributed to the shadowing effect of solid particles or improper mixing of liquid food (Marugán et al., 2008).

Log-Linear Models
Log-Linear model has been widely accepted and used to predict UV-C inactivation kinetics. This model assumes that death of micro-
organisms follows the first order kinetics. Number of microorganisms exposed to UV-C light are logarithmically reduced with expo-
sure time (Eq. 1) (Van Boekel, 2002);
   
N N t
ln ¼  kt or log ¼ (1)
NO No D
4 Kinetic and Process Modeling of UV-C Irradiation of Foods

Figure 1 Representation of typical survival curves of microorganisms. Reprinted with permission from Geeraerd, A.H., Valdramidis, V.P., and Van
Impe, J.F., 2005. GInaFiT, a freeware tool to assess non-log-linear microbial survivor curves. Int. J. Food Microbiol. 102(1), 95–105.

where N0 is the initial microbial load, N is the concentration of microorganism at a specific time (CFU/mL), k is the inactivation rate
constant (time1), t is the exposure time. D is the decimal reduction time; time required for one log reduction in number of cells and
it is equal to 2.303/k. D value is calculated from the reciprocal of the first-order rate constant commonly used to characterize the UV-
C sensitivity of microorganisms.
Delayed-Chick-Watson model (Eq. 2) with a lag time parameter (tlag) can be used for modeling the linear inactivation curve with
a shoulder (Dalrymple et al., 2010).
 
N 0 for t < tlag
log ¼   (2)
N0 k t  tlag for t > tlag

The modified Chick-Watson model (Eq. 3) can be also applied to survival data of microorganisms exposed to UV-C irradi-
ation. This model can exhibit a log linear behavior with shoulder and/or tailing (Marugán et al., 2008; Unluturk and Atilgan,
2014);
N
log ¼ k½ð1  expðk1 tÞ (3)
N0
Where k and k1 are the inactivation rate constants for linear part and tailing or shoulder region, respectively.
Geeraerd et al. (2000) suggested log-linear plus shoulder model (Eq. 4), for the survival curves exhibiting only shoulder (upward
concavity):
  SI 
ekmax
Nt ¼ N0  ekmax t
 SI
(4)
1 þ ðekmax  1Þ ekmax t

where N0 and Nt represent number of survival counts before the treatments and after time t, SI refers to shoulder length time before
exponential inactivation begins and kmax (time1) is the inactivation rate constant of exponential region.
Log linear plus tail model is also developed by Geeraerd et al. (2000) to be used in the case of inactivation curve having only tail
region (downward concavity) (Eq. 5);

Nt ¼ ðN0  Nres Þ  ekmax t


þ Nres (5)
where N0 and Nt are the number of survival microbial count before the treatments, and after time t, respectively and kinetic
parameters Nres that is the residual number of subpopulations at the end of the treatment and kmax inactivation rate of exponential
region.

Weibull Model
Inactivation curve can have both shoulder and tailing region (upward and downward concavity). In such conditions, non-linear
model described by Weibull distribution (Eq. 6) could be used (Mafart et al., 2002; Mañas and Pagán, 2005; Peleg and Penchina,
2000; Van Boekel, 2002). There are two kinetic parameters in Weibull model; p is called shape parameter and describes downward
concavity (p > 1) or upward concavity (p < 1) of the curve (if p ¼ 1, curve is linear). d called scale parameter and describes the first
decimal reduction time of survival cells (Mafart et al., 2002). Parameters of d is defined as tR, which is the reliable time, in some
studies.

N
¼ 10ðdÞ
t p
(6)
N0
Kinetic and Process Modeling of UV-C Irradiation of Foods 5

Hom Model
Hom model (Eq. 7) also used for the modeling of the inactivation curves exhibiting tailing or shoulder effect (Hom, 1972). This
model is similar to linear Chick-Watson model. Only h parameter is different. h refers to the penetration rate constant. If h is equal to
1, equation is converted to linear Chick-Watson model; in the case of h > 1 or h < 1, the curve may exhibit shoulder or tailing,
respectively (Marugán et al., 2008; Unluturk and Atilgan, 2014).
N
log ¼  kt h (7)
N0
Hom model is also modified in order to describe the sigmoidal inactivation curves having both shoulder and tailing regions
(Eq. 8). Three inactivation rate constants (k, k1, k2) in the modified-Hom model are obtained from three regions of the curve, which
are linear, shoulder and tailing regions. The modified-Hom model lets us to predict the sigmoidal shape curves.

N
log ¼ k½ð1  expðk1 tÞk2 (8)
N0

Log Logistic Model


Log-logistic model is one of the nonlinear model frequently used to describe sigmoidal shape of survival curves (Eq. 9). It was devel-
oped by Cole et al. (1993) and modified to simpler form by Chen and Hoover (2003).
N A A
log ¼  (9)
N0 1 þ e4sðslogtÞ=A 1 þ e4sðsþ6Þ=A
Where A is upper asymptotedlower asymptote (log10 CFU/mL), s is the maximum rate of inactivation [log (CFU/mL)/log min], s
is the log10 time to the maximum rate of inactivation (log10 min).

Gompertz Model
This nonlinear equation is initially proposed to model growth curves and then modified for the evaluation of microbial inactivation
kinetics (Eq. 10) (Chen and Hoover, 2003; Linton et al., 1995). The model can be used to describe the sigmoidal shape inactivation
curves observed in the framework of thermal and non-thermal inactivation.
N BðtMÞ
¼ C  ee  C  ee
BM
log (10)
N0
where C is difference between upper and lower asymptote, M is the time (min) at maximum inactivation rate, B is relative death rate
at M time.

Biphasic Model
Cerf (1977) proposed a two-fraction model in the case of two types of microorganisms or two strains of the same microorganism
exist in a food (Eq. 11). In this model, one cell is more resistant to the UV-C process than the other. In this model, f is the fraction of
initial population in a more sensitive (major) population, (1-f) is the fraction of the initial population in a more resistant (minor)
population, kmax1 and kmax2 are inactivation rate constants for more sensitive and more resistant cells, respectively (Ferrario et al.,
2013; Geeraerd et al., 2005).
N 
log ¼ log f  ekmax1 t þ ð1  f Þ  ekmax2 t (11)
N0

Baranyi Model
This model has been originally developed by Baranyi and Roberts (1994) to model the growth curve of microorganisms, and then
modified as in Eq. (12) to describe for describing nonlinear survival curves of microorganisms. In this model, shoulder, tail or both
are described with a (t) and b (t) functions (Xiong et al., 1999).
8 9
> dN
< ¼ kmax aðtÞN bðtÞ; N > 0; t  0 > = rn
dt aðtÞ ¼ 1  n
>
: >
; r þ tn
Nð0Þ ¼ N0 ; N > 0; t ¼ 0 (12)
Nmin
bðtÞ ¼ 1 
N
6 Kinetic and Process Modeling of UV-C Irradiation of Foods

where kmax is the maximum inactivation rate, r is the time required for the relative inactivation rate to reach half of kmax, n is
curvature parameter, Nmin is the minimum number of microorganism in the tailing (Nmin ¼ 0, no tailing).

Kinetic Power Law Model


The kinetic power model assumes that the inactivation kinetics is not first with respect to microbial concentration (Eq. 13). The
model can fit observed data displaying shoulders (x < 1) or tailing off behavior (x > 1) (Dalrymple et al., 2010).
N 1

ln ¼ ln 1 þ ðx  1Þkcn tN0xþ1 (13)
N0 x  1
Where c and n are constants, x is the order of reaction.
Model performance is generally evaluated using statistical measures (sum of squared estimate of errors (SSE), root mean square
error (RMSE), coefficient of determination (R2). Additionally, the model reliability is quantitatively measured by means of two
indices, bias and accuracy factor. Bias and accuracy factor are the mean square differences between predictions and observations.
They are used for validation of predictive models (Baranyi et al., 1999).

Modeling UV-C Induced Inactivation of Microorganisms on Food Surfaces

UV-C irradiation can be used to treat the surfaces of solid foods such as fresh fruits, vegetables, meat. In this section, the mathemat-
ical modeling of UV-C inactivation kinetics of target microorganisms is thoroughly discussed. Significant difference is found
between UV-C irradiation of liquid medium and the surface of a solid object. Complex interactions may occur between microor-
ganisms on a solid surface and the solid material. Especially, ‘‘shielding’’ from incident UV-C light is a major factor affecting the
disinfection performance of UV-C irradiation of the solids with different surface structure or topography (Gardner and Shama,
2000).
The published kinetic studies on UV-C surface disinfection are summarized in Table 1. The different survival curves were well
described in terms of the Weibull distribution. This model led to a better explanation about the influence the UV-C irradiation on
pathogenic and spoilage bacteria, yeasts and molds in fresh fruits (Adhikari et al., 2015; Syamaladevi et al., 2013; Syamaladevi et al.,
2014; Yun et al., 2013) and vegetables (Martinez-Hernández et al., 2015), and meat products (Chun et al., 2009). Yun et al. (2013)
investigated the efficiency of ultraviolet-C (UV-C) light for inactivation of 4–5 individual strains of Escherichia coli O157:H7 and
Salmonella spp. on apricots, in addition to a four-strain composite of Shiga toxin-negative E. coli O157:H7 and a cocktail of three
attenuated strains of Salmonella Typhimurium and Salmonella Typhimurium LT2. They stated that UV-C sensitive cells or cells outside
the fruit were rapidly reduced at the initial stage of UV-C treatment, and then, the inactivation was slowed down due to the presence
of more resistant or protected cells in the fruit. Thus, survival curves following UV-C treatment were not linear and well described by
Weibull model. Similarly, Syamaladevi et al. (2014) found Weibull model as the best model comparing to zero-order, first-order
and second-order kinetic models for the inactivation of Penicillium expansum molds in pear by UV-C irradiation. Additionally, in
apple, pear and strawberry fruits exposed to UV-C light, the concave shape (tailing) of the inactivation curve of E. coli and
L. monocytogenes was well explained by the Weibull model (Adhikari et al., 2015). Gunduz et al. (2015) investigated the germicidal
effects of ultraviolet-C (UV-C) light on the postharvest wound pathogens of citrus fruits namely Penicillium digitatum and Penicillium
italicum. Sharp initial decline and tailing was observed in the survivor curves, i.e., a subpopulation of more resistant spores declined
at a slower rate than the majority of the spores with increasing UV-C dose. They applied log linear plus tail, Baranyi and Logistic
models to experimental data and found that Logistic model has better explained the inactivation kinetics. Hakguder and Unluturk,
(2018) studied the inactivation of natural flora found on cv. ‘Şalak’ apricot surfaces using a newly designed UV-C treatment system
equipped with four UV-C lamps and a rotating roller bearing applied. The inactivation data obtained for TAPC (total mesophilic
aerobic bacteria) and YMC (yeast and mold count) were modeled by using log-linear regression, Biphasic model, log-linear þ tail,
Weibull model, Weibull þ tail, double Weibull and biphasic þ shoulder models. The inactivation kinetics of TAPC and YMC were
best described by log linear þ tail model with the lowest root mean square error (RMSE) and the highest regression coefficient
(R2 > 0.90).
Additionally, the UV-C inactivation of microorganisms such as E. coli, L. monocytogenes, Salmonella, Campylobacter on the surfaces
of fresh cut broccoli and ready to eat ham slices was best described by Weibull model (Chun et al., 2009; Martinez-Hernández et al.,
2015).

Modeling UV-C Induced Inactivation of Microorganisms in Liquid Foods

The microbiological quality of liquid food products subjected to several nonthermal processing technologies, including UV-C irra-
diation, was reviewed in detail with a focus on the mechanism of the action and the consecutive microbial reductions in different
types of juices (Roobab et al., 2018). In this section, recent advances during the last decade and the future prospects in the math-
ematical modeling of the inactivation kinetics of target microorganisms subjected to UV-C irradiation were highlighted for liquid
foods.
Table 1 Kinetic Models for UV-induced inactivation of microorganisms on food surfaces

Food UV conditions Target (Inactivation) Model D or d values References


2
Apricot UV chamber Cocktails of Salmonella (1.9 log) Weibull 18–29 mJ.cm (E.coli) Yun et al. (2013)
2 UV-C emitting bulbs (55 W) E. coli O157:H7 (1.8 log) 17–22 mJ.cm2 (Salmonella)
0–442 mJ.cm2
Ready to eat ham Batch system L. monocytogenes (4.27 log) First-order 477 J.m2 Chun et al. (2009)
UV-C lamps (15W) S. Typhimurium (4.64 log) Weibull
0–8000 J.m2 C. jejuni (5.24 log)
Broccoli Bench top UV system E. coli (1.2 log) First-order -(not calculated) Martinez-Hernández et al. (2015)
15 UV-C lamps (36W) S. Enteritidis (3.29 log) Weibull
0–15 kJ.m2 L. monocytogenes (1.14 log)
Pear UV-C EmitterTM table-top UV System Penicillium expansum (2.8 log) Zero-order 0.23 kJ.m2 (intact pear) Syamaladevi et al. (2014)
2 UV-C Emitters First-order 0.24 kJ.m2 (wounded pear)
3.1 kJ.m2 Second-order
Weibull
Pear UV-C EmitterTM table-top UV System E. coli (up to 3.6 log) Weibull 0.27 kJ.m2 (intact pear) Syamaladevi et al. (2013)
4 UV-C Emitters 0.35 kJ.m2 (wounded pear)

Kinetic and Process Modeling of UV-C Irradiation of Foods


7.56 kJ.m2 0.37 kJ.m2 (intact peach)
Orange UV-C chamber P. digitatum (3.3 log) Log-linear plus tail (GInaFit) D1:0.09 D2:24.5min (P. digitatum) Gunduz et al. (2015)
4 UV-C lamps (15W) P. italicum (2.5 log) Baranyi D1:0.41 D2:15.88 min (P. italicum)
0.26–15.84 kJ.m2 Logistic
Apricot Bench-top rotating roller UV systemTotal mesophilic aerobic bacteria (3 log) Log-linear -(not calculated) Taze and Unluturk (2018)
4 UV-C lamps (15W) Yeast-mold (2.4log) Biphasic
0–48.45 kJ.m2 Total coliform (4.7 log-completely) Log linear þ tail,
Weibull
Weibull þ tail,
Double Weibull
Biphasic þ shoulder
(GInaFiT tool)
Apple UV-C EmitterTM table-top UV System E. coli (up to 2.9 log) Weibull 0.02 kJ.m2 apple Adhikari et al. (2015)
Pear 2 UV-C Emitters L. monocytogenes (up to 1.7 log) 0.12 kJ.m2 pear
Strawberry 0.14–11.87 kJ.m2 1.69 kJ.m2 strawberry (E. coli)
0.90 kJ.m2 apple
1.35 kJ.m2 pear
40.6 kJ.m2 strawberry (L. monocytogenes)

7
8 Kinetic and Process Modeling of UV-C Irradiation of Foods

UV-C irradiation has been applied as an alternative treatment for different types of liquid foods such as apple juice (Caminiti
et al., 2012; Muller et al., 2011; Muller et al., 2014), coconut water (Ochoa-Velasco et al., 2018a), grape juice (Muller et al., 2014),
liquid egg white (Unluturk et al., 2010), mango juice (Santhirasegaram et al., 2015), pomegranate juice (Pala and Toklucu, 2011),
skim milk (Gunter-Ward et al., 2018), and white grape juice (Baysal et al., 2013). Several microorganisms (e.g., Alicyclobacillus acid-
oterrestris, E. coli, L. plantarum, L. innocua, L. monocytogenes, S. aureus, S. Typhimurium, and spoilage yeasts) have been used as a target
microorganism to be inactivated in liquid food products by UV-C light; and different types of mathematical models including log
linear, log linear biphasic, log linear plus tail, Weibull, modified Gompertz, and Hom models have been used to describe the UV-C
inactivation kinetics as summarized in Table 2.
The microbial quality of food products can be increased by UV-C irradiation. For instance, Gram negative bacteria such as E. coli
species have been reported to be efficiently inactivated by UV-C irradiation to below detectable limits in several fruit juices (Franz
et al., 2009; Kaya et al., 2015; Unluturk and Atilgan, 2014) which was in accordance with the requirement of 5-log reduction
mandated by US FDA for pathogenic microorganism capable of growing in fruit juices.
A recent study by Bhullar et al. (2018) reported that the inactivation kinetics of E. coli followed log-linear model in coconut water
subjected to the UV-C irradiation (5–30 mJ cm2). Similarly, first order kinetic model best described the inactivation kinetics of
E. coli O157:H7, S. aureus, S. Enteritidis, L. plantarum caused by UV-C light in coconut water (Donsingha and Assatarakul,
2018). It has been reported that inactivation of microorganism under UV-C irradiation can also follow a sigmoidal trend with
a shoulder and/or a tail (IFT-FDA, 2000). A non-linear microbial inactivation curve is mainly consisted of shoulder, linear and
tailing regions. A shoulder indicates an initial delay of microbial inactivation at the beginning of UV-C treatment (Marugan
et al., 2008). Tailing effect, on the other hand, has been associated with the non-homogeneous irradiation due to the suspended
solid particles and/or cell aggregation in the medium (Unluturk et al., 2008; Unluturk et al., 2010). For instance, Baysal et al.
(2013) found out that the log-linear plus tail model was best described the inactivation kinetics of A. acidoterrestris in white grape
juice. Unluturk et al. (2010) applied log-linear, Weibull, Hom, and modified Chick-Watson models to evaluate the UV-C inactiva-
tion behavior of E. coli K-12, E. coli O157:H7, L. innocua in liquid egg white exposed to UV-C irradiation at UV-C intensity of
1.314 mW cm2. The authors mainly observed linear and tailing regions in the inactivation curve which was best described with
the modified Chick Watson model with the smallest root mean square error (R2  0.92). Log-linear, Weibull, Hom, and modified
Chick-Watson models were also applied for inactivation kinetics of E. coli K-12 and lactic acid bacteria in the study of Unluturk and
Atilgan (2014) where grape juice was subjected to UV-C dose in an annular UV-C reactor varied between 31.1 and 116.7 J mL1. It
has been demonstrated that Weibull model was best fitted to the inactivation kinetics of E. coli K-12 and lactic acid bacteria
(R2 ¼ 0.999) in this case.
Regarding inactivation of yeast, insufficient elimination can be observed after UV-C irradiation depending on the food charac-
teristics and UV-C irradiation conditions. Compared to bacteria, greater UV-C dose values have been reported for yeast cells (Sinha
and Häder, 2002; Tran and Farid, 2004). Accordingly, Unluturk and Atilgan (2014) observed greater UV-C dose value for foodborne
yeast compared to E. coli K-12 in UV-C treated grape juice. Similarly, it has been demonstrated that Saccharomyces cerevisiae showed
higher resistance (3.09–4.15 log reduction) to UV-C light compared to Lactobacillus rhamnosus (7.69–9.19 log reduction) and
S. Typhimurium (6.30–9.71 log reduction) (Ochoa-Velasco et al., 2018a). Lower inactivation of yeasts compared to bacteria has
been explained by easy adaptation of yeasts to the environment with a more efficient repair system, protection of DNA by the
nuclear membrane in yeasts, lower amounts of pyrimidine in DNA of yeasts compared to bacteria (Torkamani and Niakousari,
2011). A recent study by Feliciano et al. (2019a,b) revealed a broad perspective by demonstrating the inactivation kinetic param-
eters of 17 yeasts in UV-C irradiated orange juice. The authors identified Cryptococcus albidus and Candida parapsilosis strains as the
most and the least resistant yeasts in orange juice according to the both log-linear and biphasic models with R2 > 0.90. The neces-
sary UV-C dose to achieve a 90% reduction of a microbial population (DUV-C) was reported to be 2174.63 mJ cm2 for C. albidus;
whereas S. cerevisiae (BFE-39) required DUV-C of 408.65 mJ cm2 which was not significantly different than that of the least resistant
yeast specie (i.e.C. parapsilosis, 357.88 mJ cm2) in orange juice as calculated by Baranyi and Roberts model fitting.
Different types of inactivation kinetics can be observed for microorganisms subjected to UV-C irradiation. According to the study
of Gabriel et al. (2018a), C. parapsilosis exhibited a biphasic inactivation behavior in a calamansi juice with an initial lag phase and
subsequent log-linear phase. Likewise, UV-C inactivation curve of several yeast strains in orange juice exhibited an initial lag phase
where minimal cells were inactivated (Feliciano et al., 2019a,b). Gouma et al. (2015b) also observed initial lag phases for UV-C
treated S. cerevisiae, Saccharomyces bayanus, Zygosaccharomyces bailii, Dekkera bruxellensis, and Dekkera anomala in McIlvane buffer.
As described by Baranyi and Roberts (1994), the biphasic inactivation pattern composes of an initial inactivation lag phase (or
shoulder) where minimal cells are inactivated, and follows by a fast logarithmic-linear inactivation phase. An inactivation curve
with downward concavity or a shoulder implies that damage accumulation lowers the resistance of the cell (Peleg, 2006). In the
case of biphasic curves, the exposure time necessary to reduce the population of tested microorganism should be calculated by
summing up the lag time and D UV-C value obtained from the log-linear curve. In this respect, DUV-C for big (BC1) and small
(SC1) colonies of C. parapsilosis were reported as 126.24 and 183.85 mJ cm2 in UV-C treated calamansi juice, respectively (Gabriel
et al., 2018a). Additionally, Unluturk and Atilgan (2014) demonstrated that modified Chick-Watson model best described the UV-
C inactivation kinetics of foodborne yeasts with R2 ¼ 0.999 and RMSE ¼ 0.001.
Several studies indicated that turbulent flow provides more uniform irradiation that can increase the effectiveness of UV-C light
and results in high microbial reductions (Keyser et al., 2008; Koutchma, 2008). Furthermore, use of UV-C light in combination with
other technologies is of great interest to increase the microbial inactivation and to increase the applicability of UV-C light to
different types of food products.
Table 2 Kinetic Models for UV-induced inactivation of microorganisms in liquid foods

Food UV conditions Target microorganism Model D or d values References

Coconut water Continuous flow • Two surrogate viruses (MS2, T1UV), • Log-linear • 2.82–4.54 mJ cm2 Bhullar et al. (2018)
40 W UV-C lamp • E. coli ATCC 25922,
Flow rate 36–215 mL$min1 • S. Typhimurium ATCC 13311,
UV intensity: 5–30 mJ cm2 • L. monocytogenes ATCC 19115.
Coconut water Continuous flow • E. coli O157:H7, • Zero order • Not calculated Donsingha and
8 W UV-C lamps • S. aureus, • First order Assatarakul (2018)
UV dose: 0.2–12.0 J mL1 • S. Enteritidis,
• L. plantarum
Coconut liquid endosperm Batch • Spoilage yeasts (Candida spp., Pichia spp., • Log-linear • 122.72 (linear) and 214.89 (biphasic) Feliciano et al. (2019a)
beverage 15-W UV-C lamps D. hansenii, S. cerevisiae, T. delbrueckii, • Biphasic mJ$cm2 for C. albidus
UV intensity: 3.42–4.99 mW$cm2 lamp-to- • 17.34 (linear) and 17.35 (biphasic)

Kinetic and Process Modeling of UV-C Irradiation of Foods


C. lusitaniae, C. albidus, K. marxianus,
sample surface distance: 100 mm. M. guilliermondii, T. cutaneum) mJ$cm2 for T. delbrueckii
Initial inoculum: 4–5 log CFU$mL1.
Sample thickness: 7 mm
Temperature: 25  C
Time: 0–90 s
Coconut water Flow rate: 0.5–8.0 mL s1 • L. rhamnosus, • Beta function • Not calculated Ochoa-Velasco et al.
Time: 0–10 min • S. Typhimurium, (2018)
UV-dose: 5.28–26.4 J cm2, • S. cerevisiae
Apple juice UV-C intensity (1.31, 0.71 and • A. acidoterrestris (DSM 3922 spores) • Log-linear plus • 9.55  3.54 mJ cm2 for white grape Baysal et al. (2013)
White grape juice 0.38 mW cm2) exposure time tail juice
(3–15 min) • Weibull • Not calculated for apple juice
Constant depth (0.15 cm)
UV-doses: 0–489 mJ cm2 for grape fruit
juice and 0–539 mJ cm2 for apple juice
Orange juice Batch • 17 spoilage yeasts • Linear • 1924.31 and 2174.63 mJ cm2 for Feliciano et al. (2019b)
15-W UV-C lamps • Biphasic Cryptococcus albidus
UV-C intensity: 3.64–4.97 mW$cm2 lamp- • 245.83 and 357.88 mJ cm2 for
to-sample surface distance: 100 mm. Candida parapsilosis
Inoculum: 4–5 log CFU/mL. sample
thickness: 7 mm
Temperature: 25  C
Time: 0–130 s
(Continued)

9
10
Table 2 Kinetic Models for UV-induced inactivation of microorganisms in liquid foodsdcont'd

Food UV conditions Target microorganism Model D or d values References

Kinetic and Process Modeling of UV-C Irradiation of Foods


Calamansi juice drink Batch • C. parapsilosis (BC1 and SC1 isolates) • Biphasic • 126.24 mJ cm2 (BC1) Gabriel et al. (2018)
UV-C intensity • 183.85 mJ cm2 (SC1)
15 W UV-C lamps
Film thickness: 7 mm lamp-to-sample
surface distance: 10 cm
Time: 0–50 s
Grape juice Annular flow • E. coli K-12, • Log-linear • Calculated in terms of time Unluturk and Atilgan
UV dose: 31.1–116.7 J mL1 • Lactic acid bacteria (LAB), • Weibull (2014)
Flow rate: 0.90–3.70 mL s1 • Foodborne yeasts • Hom
Circulation: five times • Modified Chick-
Watson
Liquid egg white Bench • E. coli K-12, • Log-linear • Calculated in terms of time Unluturk et al. (2010)
UV intensity: 1.314 mW cm2 • E. coli O157:H7, • Weibull
Sample thickness: 0.153 cm • L. innocua • Hom
Time: 0–20 min • Modified Chick-
Watson
Skim milk Bench top collimated beam • Two surrogate viruses (bacteriophages • Log-linear • 1.795 mJ cm2 for E.coli Gunter-Ward et al. (2018)
UV-C doses: 0–40 mJ cm2 MS2 and T1UV) • Exponential • 3.602 mJ cm2 for S. Typhimurium
• E. coli • 2.460 mJ cm2 for L. monocytogenes
• S. Typhimurium
• L. monocytogenes
Kinetic and Process Modeling of UV-C Irradiation of Foods 11

Combined UV-C Treatment With Other Technologies

Processing conditions, equipment, and treatment media characteristics can be highly influential on the microbial inactivation rates.
For instance, the lethal effect of UV-C treatments can depend on the absorptivity of the liquid media due to the fact that UV-C light is
no longer available for microbial inactivation when it is absorbed by some components of the food products (Gayan et al., 2011).
Turbulent flow pattern has been reported to prevent microbial clumping, to provide improved homogeneity of the flow, and to
contribute to the efficiency of UV-C irradiation (Keyser2008; Koutchma, 2008). When a single technique fails to achieve the
intended reduction of pathogenic bacteria in food, technologies/treatments with different bacterial inactivation mechanisms are
combined so that synergistic effects can be achieved (Fan et al., 2017). In order to increase the germicidal effect, use of UV-C light
in combination with other technologies is a promising strategy that can improve the microbial stability of especially liquid food
products. Combined treatment of UV-C irradiation and heating is one of the hurdle approaches that have been widely applied
by many researchers for the inactivation of E. coli (Carrillo et al., 2017; Gabriel et al., 2018b; Gayan et al., 2011; Gayan et al.,
2016; Gayan et al., 2012, 2013; Gouma et al., 2015a; Palgan et al., 2011), Salmonella (Gabriel et al., 2018b; Gayan et al., 2016;
Gayan, Serrano, Raso, et al., 2012; Gouma et al., 2015a; Possas et al., 2018), L. monocytogenes (Gabriel et al., 2018b; Gayan
et al., 2016; Gouma et al., 2015a), S. aureus (Gayan et al., 2014b; Gayan et al., 2016; Gouma et al., 2015a), S. cerevisiae (Carrillo
et al., 2017; Gouma et al., 2015b) in different types of liquid products as summarized in Table 3.
The required UV-C dose to inactivate the 99.99% of the initial population of different E. coli strains varied between 11.18 J mL1
and 16.6 J mL1 in Mcllvaine buffer (Gayan et al., 2011). The combined use of UV-C irradiation and heat (UV þ H), on the other
hand, has resulted in a synergistic inactivation of E. coli in fruit juices (Gayan et al., 2012). The maximum synergistic lethal effect was
obtained at 55  C (Gayan et al., 2012) while the magnitude of synergism was previously reported to reduce at temperatures above
57.5  C (Gayan et al., 2011). A treatment of UV-C irradiation (27.10 J mL1) at 55  C for 3.58 min of treatment time has been
demonstrated to ensure 5-log reductions of the E. coli strains in apple juice without effecting its pH,  Brix, and acidity (Gayan
et al., 2013). In another study, E. coli was reported to require a UV-C dose between 13.81 J mL1 and 5.20 J mL1 in apple juice
for the temperatures between 44 and 54  C (Gouma et al., 2015a). The reduction in the population of E. coli O157:H7 due to
heat (55  C) was 4.12 log; whereas the reduction level increased to 5.94 log when combined UV-C and heat treatments were
employed (Gabriel et al., 2018b). A synergistic lethal effect was also observed for the inactivation of Salmonella enterica subsp. enter-
ica upon simultaneous application of UV-C light and heat treatments (Gayan, Serrano, Raso, et al., 2012). It has been demonstrated
that E. coli O157:H7 cocktail (DUV-C ¼ 25.26 mJ cm2) showed greater resistance to UV-C irradiation than S. enterica (DUV-
2 2
C ¼ 24.65 mJ cm ) and L. monocytogenes (DUV-C ¼ 17.30 mJ cm ) cocktails (Gabriel et al., 2018b). On the other hand (Carrillo
et al., 2017), demonstrated that E. coli showed more sensitivity to UV-C assisted heat treatment than S. cerevisiae. This was in agree-
ment with the study of Keyser et al. (2008) where it is indicated that Gram-negative bacteria are more sensitive to UV-C irradiation
than the yeasts in apple juice. Recently, Kaya and Unluturk (2019) demonstrated that UV-C treatment alone (2.30 J mL1) achieved
0.54 log CFU$mL1 reduction of S. cerevisiae while UV þ H combined treatment resulted in 5.16 log CFU$mL1 reduction at
1.01 J mL1 and 51  C. Thus, UV-C irradiation in combination with heat can be used to improve the lethal effect in liquid foods.
Besides, natural additives have also been used in combination with UV-C irradiation for the inactivation of Salmonella Typhimurium,
E. coli, and natural microbiota of liquid foods (Beristain-Bauza et al., 2018; Gabriel, 2015; Ochoa-Velasco et al., 2018b). It is remark-
able to report that combined ultrasound and UV-C irradiation treatment is also a promising application since greater inactivation
rates were found in both apple and orange juice (Gabriel, 2015). Further research is recommended to gain more comprehensive
knowledge about the use of UV-C irradiation in combination with other emerging thermal or nonthermal food processing
technologies.

Modeling Microbial Inactivation Kinetics of Combined UV-C and Heat Treatments

The mathematical models used for the inactivation kinetics of several microorganisms subjected to combined UV-C irradiation and
heat treatments are summarized in Table 4. Heat survival curves generally exhibits log-linear profile and consequently fits to the
model proposed by Bigelow and Esty (1920) (Gayan et al., 2016). Weibull model, on the other hand, has been reported as a suitable
model to describe the microbial inactivation behavior for UV-C treatment (Possas et al., 2018). Weibull distribution corresponds to
a concave upward curve when the shape parameter (p) < 1, a concave downward curve when p > 1 (Peleg, 2006). It was demon-
strated that inactivation curve of Salmonella Enteritidis cells could be described by Weibull model which was resulted in p > 1 indi-
cating the existence of a downward concavity with presence of a shoulder. This implies that Salmonella Enteritidis cells showed an
initial resistance to UV-C light at 4  C. However, application of UV-C irradiation at higher temperatures varied between 8 and 30  C
resulted in concave upward curves (p < 1) (Possas et al., 2018). Gompertz, Geeraerd, and Weibull models were applied by Carrillo
et al. (2017) for the evaluation of inactivation kinetics of E. coli, S. cerevisiae, and Penicillium fluorescens in carrot-orange juice blend
subjected to UV-C irradiation (0–10.6 kJ m2) at 40–50  C. Gompertz equation represents sigmoid survival curves with shoulder,
exponential, and tailing regions (Alzamora et al., 2010). An initial lag phase prior to log-linear inactivation was observed for
S. cerevisiae in apple juice subjected to combined UV-C irradiation and heat treatment Gouma et al. (2015b). Carrillo et al.
(2017) reported that single UV-C irradiation resulted in inactivation curves with a small shoulder and without tailing; whereas pres-
ence of tailing and absence of shoulder was obtained in case of UV-C combined with heat treatment (Carrillo et al., 2017). Accord-
ingly, Gouma et al. (2015a) reported that shoulder length decreases as the temperature increases up to 57.5  C, and disappears at
12
Kinetic and Process Modeling of UV-C Irradiation of Foods
Table 3 Hurdle approaches and target microorganisms used for liquid products

Hurdle approach Reference Medium Modeling E. coli Salmonella L. monocytogenes S. aureus S. cerevisiae Native microbiota

UV-C light and heat Carrillo et al. (2017) Carrot-orange juice blend * þ þ
Gabriel et al. (2018) Coconut liquid endosperm þ þ þ
Gayan et al. (2011) Mcllvaine buffer þ
Gayan, Serrano, Raso, et al. (2012) Mcllvaine buffer þ
Gayan et al. (2012) Orange juice þ
Gayan et al. (2013) Apple juice þ
Gayan et al. (2014a,b) Mcllvaine buffer þ
Gayan et al. (2016) Chicken and vegetable broth * þ þ þ þ
Apple juice
Orange juice
Gouma et al. (2015a) Apple juice * þ þ þ þ
Gouma et al. (2015b) Apple juice þ
Kaya and Unluturk (2019) Verjuice * þ
Palgan et al. (2011) Apple-cranberry juice blend þ
Possas et al. (2018) Soymilk * þ
UV-C light and radio frequency Ukuku and Geveke (2010) Apple juice þ
electric field
UV-C light and chemical additives Beristain-Bauza et al. (2018) Cocounut water * þ
Gabriel (2015) Apple juice * þ
Orange juice
Ochoa-Velasco et al. (2018) Grapefruit juice * þ
Usaga et al. (2017) Apple juice þ

þ
*Different models are used to fit experimental microbial inactivation data. Target microorganisms used in the study.
Kinetic and Process Modeling of UV-C Irradiation of Foods 13

Table 4 Kinetic models for inactivation of microorganisms in liquid foods subjected to combined treatments

Food Treatment conditions Target microorganism Model References

Apple juice Combined UV-heat: at 50–60  C • E. coli • Log-linear þ shoulder Gouma et al. (2015a)
• Salmonella Typhimurium, • Secondary model
• L.monocytogenes • Tertiary model
• S. aureus
Apple juice Combined UV-heat: at 50–60  C • S. cerevisiae • Geeraerd Gouma et al. (2015b)
Carrot-orange juice blend Combined UV-heat: 0–10.6 kJ m2 at • E. coli • Gompertz Carrillo et al. (2017)
40–50  C. • S. cerevisiae • Geeraerd
• P. fluorescens • Weibull
Soymilk Combined UV-heat: • Salmonella Enteritidis • Weibull as primary model Possas et al. (2018)
0–10 J cm2 at 4–30  C • Secondary model

60  C. On contrary, log-linear regression plus shoulder model proposed by Geeraerd et al. (2000) was applied for the inactivation
data of E. coli, Salmonella Typhimurium, L. monocytogenes, and S. aureus in apple juice (Gouma et al., 2015a). The authors also
applied secondary model to describe the temperature dependence of the kinetic parameters such as shoulder index (SI) and
kmax. The relationship between SI and temperature exhibited a sigmoid profile with a shoulder phase; whereas a concave upward
curve was observed to describe the thermo-dependence of the kmax parameter. Furthermore, tertiary models were used for the
comparison of UV-H resistance of E. coli, Salmonella Typhimurium, L. monocytogenes, and S. aureus. UV-C irradiation significantly
improved the lethality at temperatures above 50  C in apple juice (Gouma et al. 2015a). Since accurate predictions of microbial
inactivation curves obtained by applying different UV-C dose and temperature combinations are important to facilitate the favor-
able conditions that can achieve desired levels of inactivation, a secondary modeling was employed to understand the relationship
between the scale parameter of Weibull model and the processing temperature (Possas et al., 2018). The scale parameter (d) of Wei-
bull model represents the dose required to have a first tenfold reduction of the microbial population (Peleg, 2006). It has been
stated that d values decreased with the increased temperature from 8 to 18  C while d remain constant at temperatures between
18 and 30  C (Possas et al., 2018).
Synergistic interactions between different processing technologies improves the microbial inactivation while reducing the treat-
ment intensity (Ross et al., 2003). Inactivation of different bacterial foodborne pathogens (i.e., E. coli, Salmonella Typhimurium,
L. monocytogenes, and S. aureus) in different food matrices such as vegetable broth, chicken broth, apple and orange juice has
been investigated by Gayan et al. (2016) for the optimal UV-H conditions. The authors stated that UV-H synergistic lethal effect
on E. coli was approximately 28% and 27% for vegetable and chicken broth, respectively. The magnitude of synergism was reported
to be higher for apple (45.3%) and orange juice (69.4%). Thus the level of synergism varies among microorganism and food
matrices Gayan et al., 2016). The increase in the reduction level of S. cerevisiae was reported to be less than 1-log when UV-C irra-
diation (2.9 J mL1) was applied in combination with heat at temperatures up to 50  C. As the temperature increased, 5-log and
>5-log reductions were achieved by the UV-C treatment at 57.5  C and 60  C, respectively (Gouma et al., 2015b). Petin et al.
(2001) attributed the synergistic effect of UV-C irradiation and heat (45–60  C) to the impairment of microorganism’s DNA-
damage repair ability in the inactivation of S. cerevisiae. It has been reported that a synergistic effect was observed for S. cerevisiae
at 55  C and 57.5  C, and it decreased above 57.5  C due to the higher contribution of lethal effect of heat (Gouma et al., 2015b).
The lethality of combined UV-C and heat treatments for each microorganism would be different depending on the microorgan-
ism’s thermo-dependence behavior. For instance, L. monocytogenes was reported to be the most UV-H resistant microorganism at
25  C–44  C in apple juice exposed to UV-C dose of 19.24–13.74 J mL1. However, the scenario changed when the temperature
was ranged from 44 C to 54 C, and E. coli showed the highest resistance requiring UV-C dose of 13.81 J mL1 – 5.20 J mL1.
When the temperature was between 54  C and 60  C, L. monocytogenes appeared to show the highest resistance again and become
the target microorganism requiring a UV-C dose of 5.20–2.11 J mL1. Thus, the target microorganism to be inactivated by UV-H
treatments would vary with the processing temperature (Gouma et al., 2015a). In another study, it is stated that the UV-C dose
required for a 5-log reduction of S. cerevisiae in apple juice at room temperature was reduced by 24.8%–89.3% by applying the
combined treatment at 50–57.5  C (Gouma et al., 2015b). This finding remarkably indicates that the time required to achieve
the desired microbial reductions by heat treatments can be decreased when it is combined with UV-C irradiation. As stated by
Gouma et al. (2015a,b), combined treatments reduced the required processing time by 54.4%, 36.5%, and 25.8% at 55  C,
57.5  C, and 60  C, respectively. Thus, it can be concluded that UV-C irradiation in combination with mild temperatures allows
to achieve the required microbial inactivation either by lowering UV-C doses and treatment times compared to those of necessary
for UV-C alone treatments or lowering treatment times than those of necessary for heat alone treatments. In this respect, use of UV-C
irradiation with mild heating is a promising hurdle strategy to enhance the synergistic lethal effect on the target microorganism of
concern.
Besides, mathematical models were used for the evaluation of the effect UV-C irradiation in combination with antimicrobials on
either native microbiota or a target microorganism such as Salmonella Typhimurium in liquid food products (Beristain-Bauza et al.,
2018; Ochoa-Velasco et al., 2018b). Combined use of ultrasound and UV-C irradiation can be also favorable in order to achieve 5D
14 Kinetic and Process Modeling of UV-C Irradiation of Foods

process in a shorter time compared to ultrasound alone or UV-C alone treatments (Gabriel, 2015). Future studies should focus on
the use of UV-C irradiation in combination with different technologies to better understand the inactivation kinetics of target micro-
organisms and to establish desirable processing conditions that satisfy the required level of microbial reduction while maintaining
the product quality.

Mathematical Modeling of UV-C Processing of Foods Using Computational Fluid Dynamics (CFD)

Computational fluid dynamics (CFD) has been applied in food processing industry, including clean room design, refrigeration and
transportation of foods, static mixers and pipe flow (Collins and Ciofalo, 1991; Quarini, 1995; Scott, 1994, 1997). Besides, food
processing operations such as chilling, drying, baking, mixing, freezing, cooking, pasteurization and sterilization are dependent on
fluid flow. CFD allows the food engineers to understand the performance of food equipment and to improve the quality or safety of
food products (Norton and Sun, 2006; Xia and Sun, 2002; Scott and Richardson, 1997). Several studies, in which CFD was used to
determine physical phenomena, has been published in the area of food processing (Unluturk et al., 2004) including design of static
mixers (Scott, 1997), and pipe flow of food streams (Scott, 1996).
Liquid food materials have a broad range of optical and physical properties, various chemical components which impact the UV-
C light transmittance and dose delivery. Besides, hydrodynamic behavior of the flow in UV-C reactors also influence performance of
the systems. On the other hand, solid foods have complex surface structure and topography causing non uniform UV-C intensity
distribution and affecting the disinfection efficiency of UV-C irradiation. Hence, computational fluid dynamics (CFD) tools allow
the estimation of UV-C intensity distribution and prediction of the inactivation performance of UV-C systems. Determination of
UV-C dose distribution in a continuous flow UV-C system by CFD has been extensively investigated in mostly water treatment
(Table 5).
Downey et al. (1998) estimated the residence time and size distributions of salt particles inside the continuous flow UV-C reactor
using computational fluid dynamics (CFD) and studied the effect of particle number, particle size distribution and hydrodynamic
properties of UV-C reactor on UV-C dose. Sozzi and Taghipour (2006) compared the performance of the flow models by using
continuum (Eulerian) and dispersed phase (Lagrangian) approaches integrated with finite and infinite line source UV-C dose distri-
bution models in the design of continuous flow UV-C reactor having different inlet and outlet configurations (Sozzi and Taghipour,
2006). Ducoste (2005a,b) studied the flow dynamics and residence time distribution of the fluid inside the photoreactor by
applying numerical methods and compared the effect of the mechanical energy on the homogeneous distribution of fluid in the
reactor. Sozzi and Taghipour (2006) computed the velocity and time distributions of microbial particles in the wastewater disin-
fected in the vertical UV-C reactor having different inlet and outlet configurations by time averaged turbulent flow models combined
with CFD and compared the simulation result with the ones obtained from particle imaging system. Also they numerically inves-
tigated the effect of design configurations on particle size and velocity distribution. For this purpose, Sozzi and Taghipour (2006)
simulated the flow of wastewater inside a UV-C reactor designed with two different inlet-outlet configurations by applying Reynolds
averaged Navier-Stokes flow model. Besides, the UV-C dose distribution in the wastewater containing microbial particles was pre-
dicted by CFD considering body forces resulted in turbulent flow regime (Taghipour, 2004; Sozzi and Taghipour, 2006). Elyasi and
Taghipour (2010) computed momentum and continuity equations using CFD for the design of a continuous flow UV-C reactor by
considering time independent volumetric based inactivation rate in Eulerian approach.
Duran et al. (2009) examined the single phase liquid flow including mass transfer phenomena in commercial-type U-shape and
L-shape annular reactors both numerically (using FLUENTÒ 6.3.26) and experimentally. Several hydrodynamic turbulence models
of standard k–ε, realizable k– ε, Reynolds stress (RSM), and the Abe-Kondoh-Nagano (AKN-turbulence model with low Re number)
were used for different Re numbers in the range of 0 and 11000.
Wols et al. (2010) simulated an ozone system by applying several calculation methods using CFD and taking into consideration
of residence time distributions, ozone concentrations, and particle trajectories. Eulerian and Lagrangian approaches were applied on
fluid flow hydraulics of the system to determine the disinfection rate. Also the effect of flow regime in different reactor geometry was
examined. Wols et al. (2010) applied a CFD model based on a finite-element method (COMSOL3Dv3.4Ò) for prediction of the flow
behavior, by using an integrated approach to estimate residence time distributions, local velocity measurements and visualization of
particle trajectories. For this aim, Reynolds averaged Navier Stokes (RANS) equations combined with standard k–ε turbulence
model. Turbulent kinetic energy (k) and turbulent dissipation (ε) were solved by a matrix solver-PARDISOÒ (Wols et al., 2010).
Crapulli et al. (2010) designed a pilot-scale fluoropolymer tube UV-C photoreactor to disinfect MS2 and T1 bacteriophages by using
CFD. During simulation, design parameters such as flow rate, UV-C intensity rate, microbial inactivation and UV-C dose distribu-
tion were considered in a Eulerian framework in CFD solver. Results were compared biodosimetric data to predict the disinfection
efficiency and cost of operation.

Performing a CFD Analysis


There are generally three steps in solving a problem in the CFD analysis. The pre-processing step of the analysis is to formulate the
flow problem, to model the geometry, to establish the boundary and initial conditions, and to generate the grid by an appropriate
meshing software. In the processing step, a simulation strategy, and input parameters and files are first established. An input data file
is created listing the values of the input parameters consisted with the desired strategy. The files for the grid and initial flow solution
Table 5 Determination of UV Intensity Distribution using CFD Modeling Approach

Model Numerical
Author Type of UV system solution method Solver program Flow model Particle tracking UV Intensity model Predicted UV dose Measured UV dose

Downey et al. Thin film annular flow Salt tracer Finite FIDAP 7.62 Navier Stokes, k-3 Pathogenic fungus N/A N/A N/A
(1998) U-shaped UV reactor element turbulence model (DPM with
(steady state Lagrangian approach)
turbulent flow)
Chiu et al. Rectangular vertical UV Water Finite N/A Navier Stokes, k-3 Laboratory velocity Multiple Point Source 12–40 mJ.cm2 N/A
(1999) chamber (turbulent element turbulence model field measurement- Summation
flow) Random Walk Model
Wright and Horizontal S-Shaped Water Finite CFX5 Navier Stokes, k-3 DPM with Lagrangian Infinite Line Source 0–60 mJ.cm2 (steady N/A
Hargreaves thin film annular flow volume turbulence model approach state)
(2001) UV reactor (steady/ 0–128 mJ.cm2
unsteady turbulent (steady state)
flow)
Unluturk et al. Thin film annular flow Apple cider Finite FLUENT 5 Navier Stokes (Eqn. of DPM with Lagrangian Infinite Line Source, 20–90 mJ.cm2 18.8–78.8 mJ.cm2

Kinetic and Process Modeling of UV-C Irradiation of Foods


(2004) UV Reactor (steady volume continuity and approach (apple cider Multiple Point Source
state turbulent flow) motion) particles) Summation (with
user defined function)
Beer-Lambert Law
Ducoste, Liu, Thin film annular flow Water Finite PHONEICS Navier Stokes, k-3 DPM with Lagrangian 5-point Multiple Point 25–45 mJ.cm2 N/A
Linden UV reactor (steady volume turbulence model and Eulerian Source Summation,
(2005a) state, turbulent flow) approach
(microorganisms)
Ducoste, Linden, Closed Conduit Water Finite PHOENICS, Navier Stokes, k-3 DPM with Lagrangian RADLSI and UVCalc3D 40–200 mJ.cm2 (Four N/A
Rokjer and Liu UV reactor (Steady volume FLUENT turbulence model and Eulerian with Multiple lamp)
(2005b) state turbulent flow) RADLSI and UVCalc3D approach Segment Source
(microorganisms) Summation
Elyasi and Horizontal U-Shaped Water Finite FLUENT 6.2.26 Conservation of mass DPM with Eulerian A radiation model 0–150 mJ.cm2 0–100 mJ.cm3
Taghipour thin film annular flow volume and momentum approach (microbial based on emission of
(2006) UV reactor (steady particles) UV lamp,
turbulent flow) Beer Lambert Law
(Continued)

15
16
Kinetic and Process Modeling of UV-C Irradiation of Foods
Table 5 Determination of UV Intensity Distribution using CFD Modeling Approachdcont'd

Model Numerical
Author Type of UV system solution method Solver program Flow model Particle tracking UV Intensity model Predicted UV dose Measured UV dose

Sozzi and Horizontal L- and U- Water Finite FLUENT 6.1.22 Reynolds Averaged DPM with Lagrangian N/A N/A N/A
Taghipour Shaped thin film volume Navier Stokes k-3 approach (organic
(2006) annular flow UV turbulence model materials) Particle
reactors (steady image velocimetry
turbulent flow)
Liu et al. (2004) Closed conduit Water Finite UVCalc3D (for UV dose Navier Stokes with k-3, DPM with Lagrangian Multiple Point Source 10–100 mJ.cm2 (low 20–50 mJ.cm2
polychromatic UV volume simulation) DaVis6.2 k-u, Reynolds stress approach (microbial Summation lamp power) 50–
reactor (steady state, (for particle image) and two-fluid population) Digital 300 mJ.cm2 (high
turbulent flow) PHOENICS (for flow turbulence models particle image lamp power)
simulation) velocimetry
Munoz et al. Closed conduit Water Finite FLUENT 6.1 Navier Stokes, k-3 DPM and Discrete Multiple Segment 49.2 mJ.cm2 N/A
(2007) polychromatic UV volume turbulence model Random Walk with Source Summation
reactor (steady state, Lagrangian approach
turbulent flow)
Reichl et al. Vertical S-Shaped thin Water Finite FLUENT 6.2 Conservation of mass, DPM with Random Multiple Point Source 50–200 mJ.cm2 45–80 mJ.cm2
(2006) film annular flow UV volume momentum and Walk Model (Bacillus Summation, Linear
reactor (steady species with k-3 subtilis spores) Source Integration
turbulent flow) turbulence model
Wols et al. Horizontal Water Finite COMSOL 3D v3.4 Reynolds Averaged DPM with Lagrangian Multiple Point Source 20–80 mJ.cm2 N/A
(2010) cylindricalbench top element Navier Stokes with approach (passive Summation
UV reactor (lamps k-3, turbulence tracer with potassium
perpendicular to flow model permanganate dye)
axis), steady state
turbulent flow
Chen et al. Closed conduit UV Water Finite FLUENT 6.3 Navier Stokes with k-3, DPM with Lagrangian Modified P-1 Radiation 50–550 mJ.cm2 N/A
(2011) reactor (steady state, volume k-u, Reynolds Stress approach (microbial Model (with user
turbulent flow) and two-fluid population) defined function)
turbulence models
Kinetic and Process Modeling of UV-C Irradiation of Foods 17

is generated. Then the simulation is performed by an appropriate CFD solver. As the simulation proceeds, the solution is monitored
to determine if a “converged” solution has been obtained. The last step is post processing step. Post-processing involves extracting
the desired flow properties from the computed flow field. The computed flow properties are then compared to results from analytic,
computational, or experimental studies to establish the validity of the computed results.

Simulation of UV-C Intensity Distribution on Solid Food Surfaces


UV-C disinfection is applied to solid foods as an alternative treatment method against chemical and heat treatments. Several studies
on UV-C treatment of fruits and vegetables were conducted for increasing shelf life and improving product quality (Adhikari et al.,
2015; Manzocco et al., 2011; Syamaladevi et al., 2015; Taze and Unluturk, 2018), vegetables (Martinez-Hernández et al., 2015;
Mukhopadhyay et al., 2014; Wang et al., 2019), meat products (Chun et al., 2009) or food powders (Dogu-Baykut et al., 2014;
Erdogdu and Ekiz, 2011).
Solid foods have complex surface properties influencing the effectiveness of UV-C irradiation. There are very few studies present
investigating the UV-C intensity distribution on solid food surfaces. Trivittayasil et al. (2016) simulated UV-C intensity distribution
on strawberry surface by CFD modeling to find the optimal treatment conditions. For this aim, discrete ordinate (DO) model was
used to estimate UV-C intensity in the spatial coordinates. The DO model calculates radiation intensity as a function of absorption,
scattering, reflection, and emission. The DO model accounts for semitransparent media and considers participating media (air) in
addition to radiation at the surface. Since strawberry fruit imitates near-spherical shape, radiative model was represented by the
following equation:
Z 4p
sT 4 ss
V $ ðIð!
r ;!s Þ!
s Þþ ða þ ss ÞIð!
r ;!
s Þ ¼ an2 þ Ið! r ;!
s ’Þ4ð!
s ;!
s ’ÞdU (14)
p 4p 0
where r is radial distance and s represents the unit vector for radiation direction. T is temperature (in Kelvin), s is known as Stefan
Boltzman constant and s0 is described as scattering coefficient (m1). The variable “a” is absorption coefficient of irradiated surface
with refractive index “n” and s’ refers the dummy integration variable to estimate scattering between s and s’. They also used a set of
boundary conditions defined on the physical surfaces and domain walls for complete solution of mathematical equation.
Trivittayasil et al. (2016) created two different simulation scenario featuring simple cases of radiation transfer to test the accuracy
of the simulation. The first one was created based on actual equipment while the second one was built based on an ideal condition
which only considers radiation transfer from a lamp, excluding the effect of radiation reflection by the lamp cover (Fig. 2). In the
first model (left) there is a coverage around UV-C lamp to examine reflection effect. Incident intensity was also measured by a radi-
ometer from the lamp surface to 6 different distances. By this way, intensity level, computed from DO model at the center of expo-
sure area, was compared by the radiometric measurements. In the second model (right), coverage was removed and radiometric
measurements were repeated. The aim of this model was to validate mesh sensitivity of simulation without considering the reflec-
tion of UV-C light. The intensity of points parallel to the lamp length (Ek) was estimated (2016) using Eq. (15).
0 1
B zk zLk C
Ek ¼ dIq h@ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi þ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffiA (15)
h h þ zk h h þ zLk
2 2 2 2 2 2

where “z” is axial distance at k and L-k points within lamp length, “L”. “h” is the height from lamp surface to irradiated point. The
lamp configuration and distances between a random irradiation point and lamp was shown in Fig. 3.
Simulation of UV-C dose distribution on the surface of strawberries was carried out to find the most suitable orientation of
a strawberry when irradiated by a single lamp and to investigate the effect of the number of lamps and their positions on nine straw-
berries placed on a film tray. The geometry and shapes of strawberries were scanned by 3-D laser scanner and used for generation of
surface mesh (Fig. 4).
Different UV-C light source configurations including various number of lamps were used for the simulations. Geometries were
created by ANSYS Workbench 14.0 software (ANSYS Inc., USA) and the simulation was done by Fluent 14.0 software (ANSYS Inc.,
USA), employing a DO radiation model. Geometry of strawberries were meshed automatically and manually to investigate meshing
performance. SIMPLEÒ algorithm was used for steady state equations in addition to first order upwind method for DO model to
solve spatial discretization.
As a result of the simulation, it was determined that the first model considering the reflection of UV-C light was over-predicted
the UV-C intensity than that of measured values. Hence, a correction factor, called nominal efficiency ratio, was used to enhance the
accuracy of the calculated radiation intensity. This constant was obtained by the influence of shelf life and real emission of light
source. The multiplication of nominal efficiency ratio and actual power of the lamp provided modified lamp power. Therefore, pre-
dicted intensity was re-estimated and the simulation result was found to fit well with the measurements (Fig. 5A). In other words,
predicted incident radiation intensities matched perfectly to calculated values obtained from Eq. (15) (Fig. 5B). The effect of the
number of mesh elements on the numerical results was also investigated. The mesh quality produced by the automatic meshing
method was found to be adequate for estimation of incident radiation intensities at different points along the middle line
(Fig. 5C). Overall, this study examined and simulated well the effect of different treatment conditions on radiation intensity distri-
bution on strawberry surfaces using a CFD based radiation model applied to actual equipment and specific target samples.
18 Kinetic and Process Modeling of UV-C Irradiation of Foods

Figure 2 Schematic representation of two cases used for testing of UV intensity model. Reprinted with permission from Trivittayasil, V., Tanaka, F.,
and Uchino T., 2016. Simulation of UV-C intensity distribution and inactivation of mold spores on strawberries. Food Sci. Technol. Res. 22(2),
185–192.

Figure 3 Estimation of the UV intensity (Ek) at a random point, parallel to lamp surface. Reprinted with permission from Trivittayasil, V., Tanaka, F.,
and Uchino, T., 2016. Simulation of UV-C intensity distribution and inactivation of mold spores on strawberries. Food Sci. Technol. Res. 22(2),
185–192.

Figure 4 Configuration of nine meshed strawberries for simulation. Reprinted with permission from Trivittayasil, V., Tanaka, F., and Uchino, T.,
2016. Simulation of UV-C intensity distribution and inactivation of mold spores on strawberries. Food Sci. Technol. Res. 22(2), 185–192.

Taze and Unluturk (2018) investigated the inactivation of natural flora found on cv. ‘Şalak’ apricot surfaces using a newly
designed UV-C treatment system equipped with four UV-C lamps and a rotating roller bearing. They employed Multiple Point
Source Summation (MPSS) Model to estimate incident light intensity distribution along the lamp axis and through the radial direc-
tion. They found that UV-C intensity through the radial direction changed from 171.94 W m2 to 16.53 W m2 at the surface of the
rotating roller bearing. At radial distance of 5.7 cm, the maximum and minimum light intensity was calculated as 55.55 W m2
(axial distance is 25 cm) and 30.70 W m2 (at the both ends of the tube) respectively. Considering the light intensity distribution
along the lamp axis, it can be concluded that apricots which were placed near to the both ends of the rotating roller bearing received
lower UV-C doses. Although MPSS model over-predicted the UV-C intensity data, light intensity profile showed a good agreement
with the radiometric measurements.

Simulation of UV-C Intensity Distribution in Liquid Foods


CFD combines mathematical physics of rheological and physicochemical properties of liquid food materials, iterative methods of
numerical solutions of transport equations (heat, mass and momentum) (Norton and Sun, 2006). Few studies are available on CFD
simulation of UV-C intensity distribution through a continuous flow UV-C reactors. In the following section, the simulation of UV-
C intensity distribution in liquid foods is discussed in several types of annular flow UV-C reactors. UV-C system with a single lamp
configuration was usually simulated in those studies (Atilgan, 2013; Sozzi and Taghipour, 2005, 2006; Unluturk et at. 2004). For
example, Unluturk et al. (2004) simulated UV-C intensity distribution in a thin-film UV-C reactor (Cider Sure 1500) used for treat-
ment of apple cider (Fig. 6). Discrete Phase Model (DPM) was integrated to identify flow pattern and residence time distribution of
Kinetic and Process Modeling of UV-C Irradiation of Foods 19

Figure 5 (A) Measured and predicted radiation intensity of the center point of model with lamp cover. (B) Predicted and calculated UV intensity
values with respect to axial distance of middle line (C) Effect of mesh size by mesh sensitivity test. Reprinted with permission from Trivittayasil, V.,
Tanaka, F., and Uchino, T., 2016. Simulation of UV-C intensity distribution and inactivation of mold spores on strawberries. Food Sci. Technol. Res.
22(2), 185–192.

Figure 6 Schematic diagram of dimensions of CiderSure 1500 UV reactor. Reprinted with permission from Unluturk, S.K., Arastoopour, H., and
Koutchma, T., 2004. Modeling of UV dose distribution in a thin-film UV reactor for processing of apple cider. J. Food Eng. 65, 125–136.
20 Kinetic and Process Modeling of UV-C Irradiation of Foods

particle phase. Besides this model, UV-C intensity model was used to simulate processing of apple cider. Fluent 5 CFD code was
used to solve all the governing equations.
The flow was assumed as steady state and three-dimensional. In order to predict the particle trajectories in the liquid phase, the
Lagrangian approach was used. In this approach, the motion of each particle is governed by the individual particle dynamic equa-
tion (Eq. 16). The momentum transfer from the dispersed phase to the continuous phase is obtained by computing the change in
the momentum when an individual particle passes through each control volume.
!
d!
u p 1 CD Rep ! !  r
¼ u þ up þ g 1þ c (16)
dt sy 24 rp

Where CD is drag coefficient. sy is momentum response time. Rep is defined as particle Reynolds number. u and up are fluid and
particle velocity vectors. rc and rp are fluid phase and particle phase densities. For the fluid phase, the equation of continuity and
motion solved numerically using appropriate wall boundary conditions. For distribution of UV-C intensity, two light intensity
models developed for annular UV-C disinfection reactors: infinite-line source model (ILS) and the finite-line source model, i.e.
multiple point source summation (MPSS) model.
Infinite-line source model (ILS): This model assumes that all light is emitted radially from the line source. The absorption of
light between the line source and the quartz surface is neglected. The model computes the intensity distribution based on Beer-
Lambert’s Law for radial geometry (Eq. 17):
I dðrIÞ
¼ ¼ al I (17)
r dr
Multiple Point Source Summation (MPSS) model was used to compute UV-C radiation intensity field as a summation
of trajected particles’ received radiation from a series of co-linear point source. The light emission occurred through all
the UV-C reactor volume (Liu et al., 2004; Sozzi and Taghipour, 2005; Unluturk et at. 2004). The UV-C intensity at
a random point in the flow domain is estimated from Eq. (18) by the summation of the fluence rate values of all point
sources:
X
n 4
   li
Il ðr; zÞ ¼ I m
exp  aq tq þ al r  rq (18)
i¼1
4pl2i r

where li refers to the distance from the current location (cm) to the ith number of point source out of m sources and estimated from
the Pisagor theorem (Eq. 19);
 
l2i ¼ r 2 þ z2i (19)

It was assumed that, absorption coefficient and thickness of quartz tube (aq and tq) were neglected due to little influence to inten-
sity distribution. Hence, final form of MPSS model was expressed in Eq. (20):
X 
n
4   li
Il ðr; zÞ ¼ exp a l r  rq (20)
i¼1
4pl2i m r

Unluturk et al. (2004) simulated two cases, no flow condition and flow condition. In Case 1, UV-C intensity distribution was
only calculated based on Eqs. (19 and 20) without considering the DPM model (no-flow condition). In Case 2, the flow in the
reactor was simulated using the CFD program by incorporating the DPM model and the UV-C intensity model (flow condition).
UV-C intensity distribution in the middle of the lamp was shown in Fig. 7 (Left) obtained in Case 1. ILS resulted in an intensity
level varied in the range of 2.93–63.5 mW cm2, having the average intensity of 19.8 mW cm2. On the other hand, UV-C intensity
was estimated 1.81 mW cm2 as minimum and 90 mW cm2 as maximum having the average of 19.7 mW cm2 by MPSS model at
selected vertical distances. In spite of higher maximum intensity level of MPSS model simulation, average values of two models were
computed as very close to each other.
In case 2, simulations were performed to calculate the UV-C dose received by the 225 mm particles injected uniformly (166 parti-
cles) from the inlet of the UV-C reactor using the CFD program. The DPM model was used to calculate particle residence time and
locations. The gravity and drag force were considered in these simulations. No-slip and Neumann boundary conditions were used
for the liquid phase. The slip boundary condition was considered for the particle phase. Both the ILS model and MPSS model were
used in the simulations to calculate the UV-C dose. UV-C intensity distribution of particles is depicted in Fig. 8. It was realized that
ILS model predicted an intensity range between 36.93 and 4.69 mW cm2 with average of 16.33 mW cm2. But in contrast, inten-
sity range was calculated in 45.07–3.0 mW cm2 with the average of 12.96 mW cm2 using MPSS model. The maximum and
minimum intensity values calculated from both models were lower than those calculated in Case 1. But the computed dose interval
was compatible with UV-C dose range of 18.8–78.8 mJ cm2, obtained from biodosimetry experiments at the same conditions
(Koutchma et al., 2004). It was also identified that 75% of 166 particles exposed to UV-C dose between the range of 30–
100 mW cm2. Unluturk et al. (2004) proposed two reasons for this situation. The first reason is that most of the particles in
annular flow were moving in a preferred region, mainly at the center of flow gap receiving less UV-C dose. The second reason
may be due to the fact that most of the particles with small residence time moves in the region where the fastest flow was occurring.
Kinetic and Process Modeling of UV-C Irradiation of Foods 21

Figure 7 Radial UV intensity distribution obtained from ILS model (Left) and MPSS model (Right). Reprinted with permission from Unluturk, S.K.,
Arastoopour, H., and Koutchma, T., 2004. Modeling of UV dose distribution in a thin-film UV reactor for processing of apple cider. J. Food Eng. 65,
125–136.

Figure 8 UV intensity distribution of particles computed by ILS (Left) and MPSS (Right). Reprinted with permission from Unluturk, S.K.,
Arastoopour H., and Koutchma, T., 2004. Modeling of UV dose distribution in a thin-film UV reactor for processing of apple cider. J. Food Eng. 65,
125–136.

Thus, they successfully simulated that there were slow-flow regions in the UV-C reactor in which particles had longer residence time
and exposed to higher UV-C irradiation. Whereas, lower residence time was computed in the fast flow region which was resulted in
inadequate UV-C exposure. Besides, it was realized that there was non-uniform dose distribution through the reactor caused by the
variation in the residence time of particles flowing through the system. During the simulation, several conditions (reflection-
refraction effect of UV-C lamps, fouling effect of fluid, absorptive properties of quartz tube, suspended particles of apple cider)
were neglected. It was concluded that these operation conditions might take into consideration to estimate more accurate UV-C
intensity and dose distributions.
Atilgan (2013) investigated UV-C intensity distribution in freshly squeezed opaque grape juice (FSOGJ) in an annular type
s-shaped continuous flow UV-C reactor equipped with single lamp using numerical analysis. The geometric properties of s-
shaped UV-C reactor are summarized in Table 6. It is assumed that grape juice was composed of both liquid and particle
phases. Microorganisms was considered to constitute particle phase. In the pre-processing step, the physical geometry of
S-shaped continuous flow UV-C reactor was created to define the flow boundaries and meshed by using ANSYS 14.0Ò, equip-
ped with Design ModelerÒ and MeshÒ (Fig. 9). In the second processing step, a commercial CFD solver, FLUENTÒ 14.0, was
used. Physical properties (viscosity and density) were introduced as an input to CFD solver. Liquid velocity, determined
experimentally, was also used as an inlet boundary condition. Continuity, momentum and mass transfer equations were
combined to simulate the flow field. In addition, Discrete Phase model (DPM) was integrated with flow models for simula-
tion of motion of the microbial particles in the flow field. An injection was created for a group of microbial particles having
certain density and mass flow rate. For numerical iteration, these models were discretized in terms of finite volume element
method and solved by FLUENTÒ 14.0. In the last step, the results were examined using the post processing features of the
FLUENTÒ 14. Graphical and visual inspection of the numerical solution was carried out by particle tracking features and
contour graphs.
22 Kinetic and Process Modeling of UV-C Irradiation of Foods

Table 6 Geometric and Physical Properties of S-shaped UV Reactor and


liquid medium

Tube material Quartz glass

Glass thickness (m) 0.002


Dout cylinder (m) 0.04
Din cylinder (m) 0.03
Dflow inlet (m) 0.004
Htube (m) 0.4
Flow Gap (m) 0.005
Aflux (m2) 5.498  104
Vflux (m3) 2.199  104
Ainlet (m2) 1.26  104
Lamp properties 254 nm, low mercury
Total Power (W) 15
Qinlet (mL.sL1) 0.90
Residence time (s.cycleL1) 244
Reinlet 36.38
Viscosity (kg.mL1.sL1) 0.021
Density (kg.mL3) 1065.3
A254 (mL1) 2384.7
Reinlet 5.82
Vinlet (m.sL1) 0.011
Temperature ( C) 20
Lamp power (W) 15
Flow gap (m) 0.005

Figure 9 Grid structure of an annular s-shaped UV reactor. Reprinted with permission from Atilgan, M.R., 2013. Design of a Continuous Flow UV
Reactor for Opaque Liquid Foods by Using Computational Fluid Dynamics (CFD) (Ph.D Thesis). Retrieved from http://openaccess.iyte.edu.tr/bitstream/
handle/11147/2966/10003627.pdf?sequence¼1&isAllowed¼y.

In order to define the points, where UV-C light is poorly penetrated, the estimation of the UV-C intensity and UV-C dose
distribution in the UV-C reactor (in mW$cm2) was essential. Prior to flow simulation, user defined function (UDF) developed
for Multiple Point Source Summation (MPSS) model was compiled to simulate the UV-C intensity distribution in the flow
volume.
It was shown that the microbial particles in the fluid flow volume distributed randomly in annular gap (Fig. 10). The
velocity profiles of particles and fluid were also examined at the middle section of the flow domain (Fig. 11). It was deter-
mined that maximum velocity of fluid (0.0136 m/s) was greater than that of particles (0.009 m s1). The velocity distribution
of fluid phase and particle phase did not exhibit a symmetrical parabolic profile in the annular region. Moreover, histogram of
residence time, shown in Fig. 12, did not exhibit normal distribution. High deviation was due to bimodal distribution of
particle residence times ranging between 175 and 300 s and 350–475 s. Mean residence time of microbial particles was esti-
mated as 241.2  77.96 s. This average time was very close to the one recorded manually (244 s) at the same flow rate
(Atilgan, 2013).
Intensity values were computed in axial and radial direction of the flow gap by using MPSS model. It was found that increasing
the radial direction away from the light source caused a decrease in UV-C intensity (Fig. 13). UV-C light was able to penetrate into
liquid up to 1 mm distance from the UV-C source. Moreover, it was predicted that UV-C light emission was symmetrical in the axial
direction of the lamp. The highest UV-C intensity was predicted as 41 mW cm2 in the region very close to the lamp surface, i.e., at
the position where r ¼ 0.015 and z ¼ 0.2.
Kinetic and Process Modeling of UV-C Irradiation of Foods 23

Figure 10 Streamlines of fluid particles (A) and microbial particle distribution in the fluid domain (top view) (B). Reprinted with permission
from Atilgan, M. R., 2013. Design of a Continuous Flow UV Reactor for Opaque Liquid Foods by Using Computational Fluid Dynamics (CFD)
(Ph.D Thesis). Retrieved from http://openaccess.iyte.edu.tr/bitstream/handle/11147/2966/10003627.pdf?sequence¼1&isAllowed¼y.

Figure 11 Velocity profiles of fluid and particles. Reprinted with permission from Atilgan, M.R., 2013. Design of a Continuous Flow UV Reactor for
Opaque Liquid Foods by Using Computational Fluid Dynamics (CFD) (Ph.D Thesis). Retrieved from http://openaccess.iyte.edu.tr/bitstream/handle/
11147/2966/10003627.pdf?sequence¼1&isAllowed¼y.
24 Kinetic and Process Modeling of UV-C Irradiation of Foods

Figure 12 Residence time distribution of particles in flow domain (A) and Histogram of residence time distribution of the simulated microbial
particles (B). Reprinted with permission from Atilgan, M.R., 2013. Design of a Continuous Flow UV Reactor for Opaque Liquid Foods by
Using Computational Fluid Dynamics (CFD) (Ph.D Thesis). Retrieved from http://openaccess.iyte.edu.tr/bitstream/handle/11147/2966/10003627.pdf?
sequence¼1&isAllowed¼y.

Figure 13 UV intensity distribution of the flow domain estimated by MPSS model. Reprinted with permission from Atilgan, M.R., 2013. Design of
a continuous flow UV reactor for opaque liquid foods by using computational fluid dynamics (CFD) (Ph.D Thesis). Retrieved from http://openaccess.
iyte.edu.tr/bitstream/handle/11147/2966/10003627.pdf?sequence¼1&isAllowed¼y.
Kinetic and Process Modeling of UV-C Irradiation of Foods 25

Figure 14 UV intensity distribution calculated from MPSS model and DPM model. Reprinted with permission from Atilgan, M.R., 2013.
Design of a Continuous Flow UV Reactor for Opaque Liquid Foods by Using Computational Fluid Dynamics (CFD) (Ph.D Thesis). Retrieved from
http://openaccess.iyte.edu.tr/bitstream/handle/11147/2966/10003627.pdf?sequence¼1&isAllowed¼y.

Figure 15 UV dose distribution of microbial particles using MPSS model and DPM model. Reprinted with permission from Atilgan, M.R., 2013.
Design of a Continuous Flow UV Reactor for Opaque Liquid Foods by Using Computational Fluid Dynamics (CFD) (Ph.D Thesis). Retrieved from
http://openaccess.iyte.edu.tr/bitstream/handle/11147/2966/10003627.pdf?sequence¼1&isAllowed¼y.

It was found that different particles exposed to different amount of UV-C light. UV-C intensity distribution was varied between
0.802 and 40.627 mW$cm2 (Fig. 14). 10% of this particles received UV-C intensity at 13–41 mW$cm2 level. 90% of these particle
had exposed to UV-C intensity below 10 mW cm2. Mean intensity received by microbial particles was estimated as 6.328 
7.429 mW cm2.
UV-C dose distribution of the microbial particles were also calculated during simulation by multiplying the residence
time of each particle obtained from DPM model with the UV-C intensity estimated from MPSS model (Fig. 15). It was
revealed that % 90 of particles received UV-C dose less than 3 J cm2 (low UV-C dose). It was found that particles passing
through the region near the outer wall of the annular flow zone were not sufficiently exposed to UV-C light compared to
those passing near the UV-C lamp surface. Particle tracking, depicted in Fig. 10, showed that microbial particles mostly
traveled far from UV-C light source. In addition, the high absorption coefficient (2384.7 m1) of the simulated opaque
liquid food caused low UV-C light penetration and low UV-C dose exposure of microbial particles passing through the
region near the outer wall.

Concluding Remarks

CFD analysis is capable of identifying the variability in dose distribution that otherwise would be impossible to identify by a bio-
dosimetry test. To enable accurate assessment of the effectiveness of UV-C systems for the treatment of solid and liquid foods, the
use of computational fluid dynamics (CFD) improves the description of dose distribution.
Mathematical modeling of UV-C intensity and dose distribution on solid food surfaces reveals that the power of light source,
distance to irradiated region, UV-C dose model and orientation of solid foods are critical factors affecting the inactivation efficiency
26 Kinetic and Process Modeling of UV-C Irradiation of Foods

of UV-C irradiation treatment. On the other hand, mathematical modeling of UV-C inactivation in annular flow through UV-C reac-
tors provide information about the location of the least treated liquid and the dead spots in the reactor.
It is recommended that numerical models used for accurate simulation of UV-C disinfection process should combine the UV-C
intensity models for UV-C light intensity distribution, flow models for flow field prediction in liquid foods, and microbial inacti-
vation kinetic models to represent the response of target microorganisms. Furthermore, the modeling of essential nutrient loss (such
as vitamins, carotenoids, phytochemicals) and enzyme inactivation needs to be considered during the simulation of UV disinfection
of foods. For this aim, theoretical analysis should be carried out and verified with experimental data obtained for nutrient loss and
enzyme inactivation. For this purpose, essential nutrient loss and enzyme activity might be measured by instrumental procedure
and data can be used for development of a mathematical model describing the kinetic degradation of nutrients and enzymes after
UV-C irradiation of food material. In another way, a model solution, which simulates the food material, can be used for evaluation
of the effect of UV-C light on several physical property changes (pH, Brix, etc.). Then, a model describing the kinetic data obtained
from these measurements can be incorporated as a user defined function in CFD simulation to investigate the effects of UV-C light
on the potential degradation of biochemical ingredients in UV-C processed foods. Combination of conservation of energy, mass
and momentum equations with kinetic models describing the inactivation of essential nutrients, enzymes and microorganisms
can be integrated within the same numerical solution algorithm.
In summary, CFD modeling allows to simulate the different UV-C processing conditions by combining the UV-C intensity distri-
bution with microbial inactivation kinetic models as well as the enzyme inactivation and essential nutrient degradation kinetics.
Thus, it helps to optimize the cost effectiveness and the efficiency of the small scale and large scale UV-C processes.

References

Adhikari, A., Syamaladevi, R.M., Killinger, K., Sablani, S.S., 2015. Ultraviolet-C light inactivation of Escherichia coli O157 : H7 and Listeria monocytogenes on organic fruit surfaces.
Int. J. Food Microbiol. 210, 136–142.
Alzamora, S.M., Guerrero, S.N., López-Malo, A., Palou, E., Char, C.D., Raffellini, S., 2010. Models for microorganism inactivation: application in food preservation design. In:
Processing Effects on Safety and Quality of Foods. CRC-Press, USA, pp. 87–115.
Atilgan, M.R., 2013. Design of a Continuous Flow UV Reactor for Opaque Liquid Foods by Using Computational Fluid Dynamics (CFD). PhD Thesis. Izmir, Turkey. Retrieved from.
http://openaccess.iyte.edu.tr/bitstream/handle/11147/2966/10003627.pdf?sequence¼1&isAllowed¼y.
Baranyi, J., Roberts, T., 1994. A dynamic approach to predicting bacterial growth in food. Int. J. Food Microbiol. 23 (3–4), 277–294.
Baranyi, J., Pin, C., Ross, T., 1999. Validating and comparing predictive models. Int. J. Food Microbiol. 48 (3), 159–166.
Barbosa-Cánovas, G.V., Tapia, M.S., Cano, M.P., 2005. Novel Food Processing Technologies. CRC Press. Retrieved from. https://www.crcpress.com/Novel-Food-Processing-
Technologies/Barbosa-Canovas-Tapia-Cano/p/book/9780824753337.
Baysal, A.H., Molva, C., Unluturk, S., 2013. UV-C light inactivation and modeling kinetics of Alicyclobacillus acidoterrestris spores in white grape and apple juices. Int. J. Food
Microbiol. 166 (3), 494–498.
Beristain-Bauza, S., Martinez-Nino, A., Ramirez-Gonzalez, A.P., Avila-Sosa, R., Ruiz-Espinosa, H., Ruiz-Lopez, I.I., Ochoa-Velasco, C.E., 2018. Inhibition of Salmonella typhimurium
growth in coconut (Cocos nucifera L.) water by hurdle technology. Food Control 92, 312–318.
Bhullar, M.S., Patras, A., Kilanzo-Nthenge, A., Pokharel, B., Yannam, S.K., Rakariyatham, K., Sasges, M., 2018. Microbial inactivation and cytotoxicity evaluation of UV irradiated
coconut water in a novel continuous flow spiral reactor. Food Res. Int. 103, 59–67.
Bialka, K.L., Demirci, A., Puri, V.M., 2008. Modeling the inactivation of Escherichia coli O157:H7 and Salmonella enterica on raspberries and strawberries resulting from exposure to
ozone or pulsed UV-light. J. Food Eng. 85, 444–449.
Bigelow, W.D., Esty, J.R., 1920. The thermal death point in relation to time of typical thermophilic organisms. J. Infect. Dis. 27, 602–617.
Bintsis, T., Litopoulou-Tzanetaki, E., Robinson, R.K., 2000. Existing and potential applications of ultraviolet light in the food industry - a critical review. J. Sci. Food Agric. 80 (6),
637–645.
Buzrul, S., Alpas, H., 2007. Modeling inactivation kinetics of foodbornepathogens at a constant temperature. LWT - Food Sci. Technol. 40 (4), 632–637.
Caminiti, I.M., Palgan, I., Munoz, A., Noci, F., Whyte, P., Morgan, D.J., Cronin, D.A., Lyng, J.G., 2012. The effect of ultraviolet light on microbial inactivation and quality attributes of
apple juice. Food Bioprocess Technol. 5 (2), 680–686.
Carrillo, M.G., Ferrario, M., Guerrero, S., 2017. Study of the inactivation of some microorganisms in turbid carrot-orange juice blend processed by ultraviolet light assisted by mild
heat treatment. J. Food Eng. 212, 213–225.
Cerf, O., 1977. A REVIEW Tailing of survival curves of bacterial spores. J. Appl. Microbiol. 42 (1), 1–19.
Char, C.D., Mitilinaki, E., Guerrero, S.N., Alzamora, S.M., 2010. Use of high-intensity ultrasound and UV-C light to inactivate some microorganisms in fruit juices. Food Bioprocess
Technol. 3 (6), 797–803.
Chen, H., Hoover, D.G., 2003. Modeling the combined effect of high hydrostatic pressure and mild heat on the inactivation kinetics of Listeria monocytogenes Scott A in whole milk.
Innov. Food Sci. Emerg. Technol. 4 (1), 25–34.
Chen, H., 2007. Use of linear, Weibull, and log-logistic functions to model pressure inactivation of seven foodborne pathogens in milk. Food Microbiol. 24 (3), 197–204.
Chen, J., Deng, B., Kim, C.N., 2011. Computational fluid dynamics (CFD) modeling of UV disinfection in a closed-conduit reactor. Chem. Eng. Sci. 66, 4983–4990.
Chiu, K., Lyn, D.A., Savoye, P., Blatchley, E.R., 1999. Integrated UV disinfection model based on particle tracking. J. Environ. Eng. 125 (1), 7--16.
Chun, H., Kim, J., Chung, K., Won, M., Bin, K., 2009. Inactivation kinetics of Listeria monocytogenes , Salmonella enterica serovar Typhimurium, and Campylobacter jejuni in ready-
to-eat sliced ham using UV-C irradiation. Meat Sci. 83 (4), 599–603.
Cole, M.B., Davies, K.W., Munro, G., Holyoak, C.D., Kilsby, D.C., 1993. A vitalistic model to describe the thermal inactivation of Listeria monocytogenes. J. Ind. Microbiol. 12 (3–5),
232–239.
Collins, M.W., Ciofalo, M., 1991. Computational fluid dynamics and its application to transport processes. J. Chem. Technol. Biotechnol. 52 (l), 5–47.
Crapulli, F., Santoro, D., Haas, C.N., Notarnicola, M., Liberti, L., 2010. Modeling virus transport and inactivation in a fluoropolymer tube UV photoreactor using computational fluid
dynamics. Chem. Eng. J. 161 (1–2), 9–18.
Dalrymple, O.K., Stefanakos, E., Trotz, M.A., Goswami, D.Y., 2010. A review of the mechanisms and modeling of photocatalytic disinfection. Appl. Catal. B Environ. 98 (1–2),
27–38.
Dogu-Baykut, E., Gurbuz, G., Decker, E.A., 2014. Impact of shortwave ultraviolet (UV-C) radiation on the antioxidant activity of thyme (Thymus vulgaris L.). Food Chem. 157,
167–173.
Donsingha, S., Assatarakul, K., 2018. Kinetics model of microbial degradation by UV radiation and shelf life of coconut water. Food Control 92, 162–168.
Kinetic and Process Modeling of UV-C Irradiation of Foods 27

Downey, D., Giles, D., Delwiche, M., 1998. Finite element analysis of particle and liquid flow through an ultraviolet reactor. Comput. Electron. Agric. 21 (2), 81–105.
Ducoste, J., Liu, D., Linden, K., 2005a. Alternative approaches to modeling fluence distribution and microbial inactivation in ultraviolet reactors: Lagrangian versus Eulerian.
J. Environ. Eng. 131 (10), 1393–1403.
Ducoste, J., Linden, K., Rokjer, D., Liu, D., 2005b. Assessment of reduction equivalent fluence bias using computational fluid dynamics. Environ. Eng. Sci. 22 (5), 615–628.
Duran, J.E., Taghipour, F., Mohseni, M., 2009. CFD modeling of mass transfer in annular reactors. Int. J. Heat Mass Transf. 52, 5390–5401.
Elyasi, S., Taghipour, F., 2006. Simulation of UV photo reactor for water disinfection in Eulerian framework. Chem. Eng. Sci. 61 (14), 4741–4749.
Elyasi, S., Taghipour, F., 2010. Simulation of UV photo reactor for degradation of chemical contaminants: model development and evaluation. Environ. Sci. Technol. 44,
2056–2063.
Erdogdu, S.B., Ekiz, H.I., 2011. Effect of ultraviolet and far infrared radiation on microbial decontamination and quality of cumin seeds. J. Food Sci. 76 (5), 284–292.
Fakruddin, M., Mazumder, R.M., Mannan, K.S.B., 2011. Predictive microbiology: modeling microbial responses in food. Ceylon J. Sci. Biol. Sci. 40 (2), 121–131.
Fan, X., Huang, R., Chen, H., 2017. Application of ultraviolet C technology for surface decontamination of fresh produce. Trends Food Sci. Technol. 70, 9–19.
Feliciano, R.J., Estilo, E.E.C., Nakano, H., Gabriel, A.A., 2019a. Decimal reduction energies of UV-C-irradiated spoilage yeasts in coconut liquid endosperm. Int. J. Food Microbiol.
290, 170–179.
Feliciano, R.J., Estilo, E.E.C., Nakano, H., Gabriel, A.A., 2019b. Ultraviolet-C resistance of selected spoilage yeasts in orange juice. Food Microbiol. 78, 73–81.
Fenner, R.A., Komvuschara, K., 2005. A new kinetic model for ultraviolet disinfection of greywater. J. Environ. Eng. 131 (6), 850–864.
Fernández, A., Lópeza, M., Bernardoa, A., Condón, S., Raso, J., 2010. Modelling thermal inactivation of Listeria monocytogenes in sucrose solutions of various water activities. Food
Microbiol. 24 (4), 372–379.
Ferrario, M., Alzamora, S., Guerrero, S., 2013. Inactivation kinetics of some microorganisms in apple, melon, orange and strawberry juices by high intensity light pulses. J. Food Eng.
118 (3), 302–311.
Franz, C.M., Specht, I., Cho, G.S., Graef, V., Stahl, M.R., 2009. UV-C-inactivation of microorganisms in naturally cloudy apple juice using novel inactivation equipment based on
Dean vortex technology. Food Control 20 (12), 1103–1107.
Gabriel, A.A., 2015. Combinations of selected physical and chemical hurdles to inactivate Escherichia coli O157:H7 in apple and orange juices. Food Control 50, 722–728.
Gabriel, A.A., Manalo, M.R., Feliciano, R.J., Garcia, N.K.A., Dollete, U.G.M., Acanto, C.N., Paler, J.C.B., 2018a. A Candida parapsilosis inactivation-based UV-C process for
calamansi (Citrus microcarpa) juice drink. LWT - Food Sci. Technol. 90, 157–163.
Gabriel, A.A., Ostonal, J.M., Cristobal, J.O., Pagal, G.A., Armada, J.V.E., 2018b. Individual and combined efficacies of mild heat and ultraviolet-c radiation against Escherichia coli
O157:H7, Salmonella enterica, and Listeria monocytogenes in coconut liquid endosperm. Int. J. Food Microbiol. 277, 64–73.
Garde-Cerdán, T., Arias-Gil, M., Marsellés-Fontanet, A.R., Ancín-Azpilicueta, C., Martín-Belloso, O., 2007. Effects of thermal and non-thermal processing treatments on fatty acids
and free amino acids of grape juice. Food Control 18 (5), 473–479.
Gardner, D.W.M., Shama, G., 2000. Modeling UV-induced inactivation of microorganisms on surfaces. J. Food Prot. 63 (1), 63–70.
Gayan, E., Condón, S., Álvarez, I., 2014a. Biological aspects in food preservation by ultraviolet light: a review. Food Bioprocess Technol. 7 (1), 1–20.
Gayan, E., Garcia-Gonzalo, D., Alvarez, I., Condon, S., 2014b. Resistance of Staphylococcus aureus to UV-C light and combined UV-heat treatments at mild temperatures. Int. J.
Food Microbiol. 172, 30–39.
Gayan, E., Monfort, S., Alvarez, I., Condon, S., 2011. UV-C inactivation of Escherichia coli at different temperatures. Innov. Food Sci. Emerg. Technol. 12 (4), 531–541.
Gayan, E., Serrano, M.J., Alvarez, I., Condon, S., 2016. Modeling optimal process conditions for UV-heat inactivation of foodborne pathogens in liquid foods. Food Microbiol. 60,
13–20.
Gayan, E., Serrano, M.J., Monfort, S., Alvarez, I., Condon, S., 2012. Combining ultraviolet light and mild temperatures for the inactivation of Escherichia coli in orange juice. J. Food
Eng. 113 (4), 598–605.
Gayan, E., Serrano, M.J., Raso, J., Alvarez, I., Condon, S., 2012. Inactivation of Salmonella enterica by UV-C Light Alone and in Combination with Mild Temperatures. Appl. Environ.
Microbiol. 78 (23), 8353–8361.
Gayan, E., Serrano, M.J., Monfort, S., Alvarez, I., Condon, S., 2013. Pasteurization of apple juice contaminated with Escherichia coli by a combined UV-mild temperature treatment.
Food Bioprocess Technol. 6 (11), 3006–3016.
Geeraerd, A.H., Herremans, C.H., Van Impe, J.F., 2000. Structural model requirements to describe microbial inactivation during a mild heat treatment. Int. J. Food Microbiol. 59 (3),
185–209.
Geeraerd, A.H., Valdramidis, V.P., Van Impe, J.F., 2005. GInaFiT, a freeware tool to assess non-log-linear microbial survivor curves. Int. J. Food Microbiol. 102 (1), 95–105.
Geveke, D.J., 2008. UV inactivation of E. coli in liquid egg white. Food Bioprocess Technol. 1 (2), 201–206.
Gouma, M., Alvarez, I., Condon, S., Gayan, E., 2015a. Modelling microbial inactivation kinetics of combined UV-H treatments in apple juice. Innov. Food Sci. Emerg. Technol. 27,
111–120.
Gouma, M., Gayan, E., Raso, J., Condon, S., Alvarez, I., 2015b. Inactivation of spoilage yeasts in apple juice by UV-C light and in combination with mild heat. Innov. Food Sci. Emerg.
Technol. 32, 146–155.
Guerrero-Beltrán, J.A., Barbosa-Cánovas, G.V., 2004. Review: advantages and limitations on processing foods by UV light. Food Sci. Technol. Int. 10 (3), 137–147.
Guerrero-Beltrán, J.A., Barbosa-Cánovas, G.V., 2005. Reduction of Saccharomyces cerevisiae, Escherichia coli and Listeria innocua in apple juice by ultraviolet light. J. Food Process
Eng 28 (5), 437–452.
Gunduz, G.T., Juneja, V.K., Pazir, F., 2015. Application of ultraviolet-C light on oranges for the inactivation of postharvest wound pathogens. Food Control 57, 9–13.
Gunter-Ward, D.M., Patras, A., Bhullar, M.S., Kilonzo-Nthenge, A., Pokharel, B., Sasges, M., 2018. Efficacy of ultraviolet (UV-C) light in reducing foodborne pathogens and model
viruses in skim milk. J. Food Process. Preserv. 42 (2).
Hijnen, W.A.M., Beerendonk, E.F., Medema, G.J., 2006. Inactivation credit of UV radiation for viruses, bacteria and protozoan (oo) cysts in water: a review. Water Res. 40 (1), 3–22.
Hom, L.W., 1972. Kinetics of chlorine disinfection in an ecosystem. J. Sanit. Eng. Div. 98 (1), 183–194.
Huang, L., 2009. Thermal inactivation of Listeria monocytogenes in ground beef under isothermal and dynamic temperature conditions. J. Food Eng. 90, 380–387.
IFT-FDA, 2000. Kinetics of Microbial Inactivation for Alternative Food Processing Technologies. IFT/FDA Contract No. 223-98-2333. Institute of Food Technologists - Food and Drug
Administration. Retrieved from. https://www.fda.gov/media/103619/download. (Accessed 29 July 2019).
Johnson, K.M., Kumar, M.R.A., Ponmurugan, P., Gananamangai, B.M., 2010. Ultraviolet radiation and its germicidal effect in drinking water purification. J. Phytol. 5 (2), 12–19.
Kaya, Z., Unluturk, S., 2019. Pasteurization of verjuice by UV-C irradiation and mild heat treatment. J. Food Process. Eng. 42 (5), e13131.
Kaya, Z., Yildiz, S., Unluturk, S., 2015. Effect of UV-C irradiation and heat treatment on the shelf life stability of a lemon-melon juice blend: multivariate statistical approach. Innov.
Food Sci. Emerg. Technol. 29, 230–239.
Keyser, M., Muller, I.A., Cilliers, F.P., Nel, W., Gouws, P.A., 2008. Ultraviolet radiation as a non-thermal treatment for the inactivation of microorganisms in fruit juice. Innov. Food
Sci. Emerg. Technol. 9 (3), 348–354.
Koutchma, T., 2008. UV light for processing foods. Ozone: Sci. Eng. 30 (1), 93–98.
Koutchma, T., 2014. Preservation and Shelf Life Extension UV Applications for Fluid Foods. Elsevier.
Koutchma, T.N., Forney, L.J., Larry, J., Moraru, C.I., 2009. Ultraviolet Light in Food Technology : Principles and Applications. CRC Press.
Koutchma, T.N., Keller, S., Chirtel, S., Parisi, B., 2004. Ultraviolet disinfection of juice products in laminar and turbulent flow reactors. Innov. Food Sci. Emerg. Technol. 5 (2),
179–189.
Linton, R.H., Carter, W., Pierson, M., Hackney, C., 1995. Use of a modified Gompertz equation to model non-linear survival curves and predict temperature, pH and sodium chloride
effects for Listeria monocytogenes Scott A. J. Food Prot. 58 (9), 946–954.
28 Kinetic and Process Modeling of UV-C Irradiation of Foods

Liu, D., Ducoste, J., Shanshan, J., Linden, K., 2004. Evaluation of alternative fluence rate distribution models. J. Water Supply Res. Technol. - Aqua 53 (6), 391–408.
Mafart, P., Couvert, O., Gaillard, S., Leguerinel, I., 2002. On calculating sterility in thermal preservation methods: application of the Weibull frequency distribution model. Int. J. Food
Microbiol. 72, 107–113.
Mañas, P., Pagán, R., 2005. Microbial inactivation by new technologies of food preservation. J. Appl. Microbiol. 98 (6), 1387–1399.
Manzocco, L., Da Pieve, S., Bertolini, A., Bartolomeoli, I., Maifrenia, M., Vianello, A., Nicoli, C.M., 2011. Surface decontamination of fresh-cut apple by UV-C light exposure: effects
on structure, colour and sensory properties. Postharvest Biol. Technol. 61, 165–171.
Martinez-Hernández, G., Huertas, J., Navarro-Rico, J., Gomez, P., Artes, F., Palop, A., Artes-Hernandes, F., 2015. Inactivation kinetics of foodborne pathogens by UV-C radiation
and its subsequent growth in fresh-cut kailan-hybrid broccoli. Food Microbiol. 46, 263–271.
Marugan, J., van Grieken, R., Sordo, C., Cruz, C., 2008. Kinetics of the photocatalytic disinfection of Escherichia coli suspensions. Appl. Catal. B Environ. 82 (1–2), 27–36.
Mukhopadhyay, S., Ukuku, D.O., Juneja, V., Fan, X., 2014. Effects of UV-C treatment on inactivation of Salmonella enterica and Escherichia coli O157:H7 on grape tomato surface
and stem scars, microbial loads, and quality. Food Control 44, 110–117.
Muller, A., Stahl, M.R., Graef, V., Franz, C.M.A.P., Huch, M., 2011. UV-C treatment of juices to inactivate microorganisms using Dean vortex technology. J. Food Eng. 107 (2),
268–275.
Muller, A., Stahl, M.R., Greiner, R., Posten, C., 2014. Performance and dose validation of a coiled tube UV-C reactor for inactivation of microorganisms in absorbing liquids. J. Food
Eng. 138, 45–52.
Munoz, A., Craik, S., Kresta, S., 2007. Computational fluid dynamics for predicting performance of ultraviolet disinfection-sensitivity to particle tracking inputs. J. Environ. Eng. Sci.
6, 285–301.
Murakami, E.G., Jackson, L., Madsen, K., Schickedanz, B., 2006. Factors affecting the ultraviolet inactivation of Escherichia coli K12 in apple juice and a model system. J. Food
Process. Eng. 29 (1), 53–71.
Ngadi, M., Smith, J.P., Cayouette, B., 2003. Kinetics of ultraviolet light inactivation of Escherichia coli O157:H7 in liquid foods. J. Sci. Food Agric. 83 (15), 1551–1555.
Norton, T., Sun, D., 2006. Computational fluid dynamics (CFD)-An effective and efficient design and analysis tool for the food industry: a review. Trends Food Sci. Technol. 17,
600–620.
Ochoa-Velasco, C.E., Diaz-Lim, M.C., Avila-Sosa, R., Ruiz-Lopez, I.I., Corona-Jimenez, E., Hernandez-Carranza, P., López-Malo, A., Guerrero-Beltran, J.A., 2018a. Effect of UV-C
light on Lactobacillus rhamnosus, Salmonella Typhimurium, and Saccharomyces cerevisiae kinetics in inoculated coconut water: survival and residual effect. J. Food Eng. 223,
255–261.
Ochoa-Velasco, C.E., Salcedo-Pedraza, C., Hernandez-Carranza, P., Guerrero-Beltran, J.A., 2018b. Use of microbial models to evaluate the effect of UV-C light and trans-
cinnamaldehyde on the native microbial load of grapefruit (Citrus x paradisi) juice. Int. J. Food Microbiol. 282, 35–41.
Pala, C.U., Toklucu, A.K., 2011. Effect of UV-C light on anthocyanin content and other quality parameters of pomegranate juice. J. Food Compos. Anal. 24 (6), 790–795.
Palgan, I., Caminiti, I.M., Munoz, A., Noci, F., Whyte, P., Morgan, D.J., Cronin, D.A., &Lyng, J.G., 2011. Combined effect of selected non-thermal technologies on Escherichia coli
and Pichia fermenters inactivation in an apple and cranberry juice blend and on product shelf life. Int. J. Food Microbiol. 151 (1), 1–6.
Peleg, M., 2006. Isothermal Microbial Heat Inactivation Advanced Quantitative Microbiology for Foods and Biosystems: Models for Predicting Growth and Inactivation. CRC Press,
pp. 1–48.
Peleg, M., Penchina, C.M., 2000. Modeling microbial survival during exposure to a lethal agent with varying intensity. Crit. Rev. Food Sci. Nutr. 40 (2), 159–172.
Pereira, R.N., Vicente, A.A., 2010. Environmental impact of novel thermal and non-thermal technologies in food processing. Food Res. Int. 43 (7), 1936–1943.
Petin, V.G., Kim, J.K., Rassokhina, A.V., Zhurakovskaya, G.P., 2001. Mitotic recombination and inactivation in Saccharomyces cerevisiae induced by UV-radiation (254 nm) and
hyperthermia depend on UV fluence rate. Mutat. Res. Fundam. Mol. Mech. Mutagen. 478 (1–2), 169–176.
Possas, A., Valero, A., Garcia-Gimeno, R.M., Perez-Rodriguez, F., de Souza, P.M., 2018. Influence of temperature on the inactivation kinetics of Salmonella enteritidis by the
application of UV-C technology in soymilk. Food Control 94, 132–139.
Quarini, J., 1995. Applications of computational fluid dynamics in food and beverage production. Food Sci. Technol. 9 (4), 234–237.
Reichl, C., Buchner, C., Hirschmann, G., Sommer, R., Cabaj, A., 2006. Development of a simulation method to predict UV disinfection reactor performance and comparison to
biodosimetric measurements. In: Conference on Modelling Fluid Flow (CMFF’06) the 13th International Conference on Fluid Flow Technologies, Budapest, Hungary.
Retamar, M.E.M., Passalía, C., Brandi, R.J., Labas, M.D., 2019. Dose estimation methodology for the UV inactivation of bioaerosols in a Continuous-Flow reactor. Aerosol Sci.
Technol. 53 (1), 8–20.
Roobab, U., Aadil, R.M., Madni, G.M., Bekhit, A.E.D., 2018. The impact of nonthermal technologies on the microbiological quality of juices: a review. Compr. Rev. Food Sci. Food
Saf. 17 (2), 437–457.
Ross, A.I., Griffiths, M.W., Mittal, G.S., Deeth, H.C., 2003. Combining nonthermal technologies to control foodborne microorganisms. Int. J. Food Microbiol. 89 (2), 125–138.
Santhirasegaram, V., Razali, Z., George, D.S., Somasundram, C., 2015. Comparison of UV-C treatment and thermal pasteurization on quality of Chokanan mango (Mangifera indica
L.) juice. Food Bioprod. Process. 94, 313–321.
Schaffner, D.W., Labuza, T.P., 1997. Predictive Microbiology: Where Are We, and Where Are We Going? Food Technology (USA). Institute of Food Technologists. Retrieved from.
http://agris.fao.org/agris-search/search.do?recordID¼US9733792.
Scott, C.M., 1994. Computational fluid dynamics for the food industry. Food Technol. Int Eur. 49–51.
Scott, G.M., 1997. Simulation of the Flow of Non-newtonian Foods Using Computational Fluid Dynamics. Campden Chorleywood Food Research Association RD Report No. 34,
Chipplng Campden, UK.
Scott, G., Richardson, P., 1997. The application of computational fluid dynamics in the food industry. Trends Food Sci. Technol. 81, 119–124.
Scott, G.M., 1996. Computational Fluid Dynamics-Modelling the Flow of Newtonian Fluids in Pipelines. Campden Chorleywood Food Research Association RD Report No. 24,
Chipping Campden, UK.
Shah, N.N.A.K., Shamsudin, R., Rahman, A.R., Adzahan, N., 2016. Fruit juice production using ultraviolet pasteurization: a review. Beverages 2 (3), 22.
Sinha, R.P., Häder, D.P., 2002. UV-induced DNA damage and repair: a review. Photochem. Photobiol. Sci. 1 (4), 225–236.
Sizer, C.E., Balasubramaniam, V.M., 1999. New intervention processes for minimally processed juices. Food Technol. Retrieved from http://agris.fao.org/agris-search/search.do?
recordID¼US2000003312.
Sozzi, D.A., Taghipour, F., 2006. UV reactor performance modeling by Eulerian and Lagrangian methods. Environ. Sci. Technol. 40, 1609–1615.
Sozzi, D.A., Taghipour, F., 2005. Experimental investigation of flow field in annular ultraviolet reactors using particle image velocimetry. Ind. Eng. Chem. Res. 44, 9979–9988.
Syamaladevi, R.M., Lu, X., Sablani, S.S., Insan, S.K., 2013. Inactivation of Escherichia coli population on fruit surfaces using ultraviolet-C light : influence of fruit surface
characteristics. Food Bioprocess Technol. 6 (11), 2959–2973.
Syamaladevi, R.M., Lupien, S.L., Bhunia, K., Sablani, S.S., Dugan, F., Rasco, B., Killinger, K., Dhingra, A., Ross, C., 2014. UV-C light inactivation kinetics of Penicillium expansum
on pear surfaces : influence on physicochemical and sensory quality during storage. Postharvest Biol. Technol. 87, 27–32.
Syamaladevi, R.M., Adhikari, A., Lupien, S.L., Dugan, F., Bhunia, K., Dhingra, A., Sablani, S.S., 2015. Ultraviolet-C light inactivation of Penicillium expansum on fruit surfaces. Food
Control 50, 297–303.
Taghipour, F., 2004. Ultraviolet and ionizing radiation for microorganism inactivation. Water Resour. 38, 3940–3948.
Taze, B.H., Unluturk, S., 2018. Effect of postharvest UV-C treatment on the microbial quality of ‘Şalak’ apricot. Sci. Hortic. 233, 370–377.
Torkamani, A.E., Niakousari, M., 2011. Impact of UV-C light on orange juice quality and shelf life. Int. J. Food Res. Technol. 18, 1265–1268.
Tran, M.T.T., Farid, M., 2004. Ultraviolet treatment of orange juice. Innov. Food Sci. Emerg. Technol. 5 (4), 495–502.
Trivittayasil, V., Tanaka, F., Uchino, T., 2016. Simulation of UV-C intensity distribution and inactivation of mold spores on strawberries. Food Sci. Technol. Res. 22 (2), 185–192.
Kinetic and Process Modeling of UV-C Irradiation of Foods 29

U.S. FDA, Irradiation in the production, processing and handling of food. Code of Federal Regulations. Title 21, part 179. Federal Register. 65: 71056–71058. U.S. Food and Drug
Administration, Washington, DC 2000.
Ukuku, D.O., Geveke, D.J., 2010. A combined treatment of UV-light and radio frequency electric field for the inactivation of Escherichia coli K-12 in apple juice. Int. J. Food
Microbiol. 138 (1–2), 50–55.
Unluturk, S.K., Arastoopour, H., Koutchma, T., 2004. Modeling of UV dose distribution in a thin-film UV reactor for processing of apple cider. J. Food Eng. 65, 125–136.
Unluturk, S., Atilgan, M.R., 2014. UV-C irradiation of freshly squeezed grape juice and modeling inactivation kinetics. J. Food Process. Eng. 37 (4), 438–449.
Unluturk, S., Atilgan, M.R., Baysal, A.H., Tari, C., 2008. Use of UV-C radiation as a non-thermal process for liquid egg products (LEP). J. Food Eng. 85 (4), 561–568.
Unluturk, S., Atilgan, M.R., Baysal, A.H., Unluturk, M.S., 2010. Modeling inactivation kinetics of liquid egg white exposed to UV-C irradiation. Int. J. Food Microbiol. 142 (3),
341–347.
Usaga, J., Manns, D.C., Moraru, C.I., Worobo, R.W., Padilla-Zakour, O.I., 2017. Ascorbic acid and selected preservatives influence effectiveness of UV treatment of apple juice.
LWT - Food Sci. Technol 75, 9–16.
Valdramidis, V.P., Bernaerts, K., Van Impe, J.F., Geeraerd, A.H., 2005. An alternative approach to non-log-linear thermal microbial inactivation: modelling the number of log cycles
reduction with respect to temperature. Food Technol. Biotechnol. 43 (4), 321–327.
Van Boekel, M.A.J.S., 2002. On the use of the Weibull model to describe thermal inactivation of microbial vegetative cells. Int. J. Food Microbiol. 74 (1–2), 139–159.
Walkling-Ribeiro, M., Noci, F., Cronin, D.A., Riener, J., Lyng, J.G., Morgan, D.J., 2008. Reduction of Staphylococcus aureus and quality changes in apple juice processed by
ultraviolet irradiation, pre-heating and pulsed electric fields. J. Food Eng. 89 (3), 267–273.
Wang, D., Chen, L., Ma, Y., Zhang, M., Zhao, Y., Zhao, X., 2019. Effect of UV-C treatment on the quality of fresh-cut lotus (Nelumbo nucifera Gaertn.) root. Food Chem. 278,
659–664.
Wols, B.A., Shao, L., Uijttewaal, W.S.J., Hofman, J.A.M.H., Rietveld, L.C., Van Dijk, J.C., 2010. Evaluation of experimental techniques to validate numerical computations of the
hydraulics inside a UV bench-scale reactor. Chem. Eng. Sci. 65 (5), 4491–4502.
Wright, N.G., Hargreaves, D.M., 2001. The use of CFD in the evaluation of UV treatment. systems. J. Hydroinf. 3 (2), 59–70.
Xia, B., Sun, D., 2002. Applications of computational fluid dynamics (CFD) in the food industry: a review. Comput. Electron. Agric. 34, 5–24.
Xiong, R., Xie, G., Edmondson, A.S., Linton, R.H., Sheard, M.A., 1999. Comparison of the Baranyi model with the modified Gompertz equation for modelling thermal inactivation of
Listeria monocytogenes Scott A. Food Microbiol. 16 (3), 269–279.
Ye, Z., Koutchma, T., 2010. Mathematical modeling and design of ultraviolet light processes for liquid foods and beverages. In: Farid, M. (Ed.), Mathematical Modelling of Food
Processing. CRC Press, pp. 575–606.
Yun, J., Yan, R., Fan, X., Gurtler, J., Phillips, J., 2013. Fate of E. coli O157:H7 , Salmonella spp . and potential surrogate bacteria on apricot fruit , following exposure to UV-C light.
Int. J. Food Microbiol. 166 (3), 356–363.

Further Reading

CFX Release, A.N.S.Y.S., 2012. Public Notice, ANSYS, Inc., Southpointe. 13.1 Technical Specifications. USA, Canonsburg, PA.

You might also like