You are on page 1of 23

Fouling and Cleaning Studies in the Food and

Beverage Industry Classified by Cleaning Type


Kylee R. Goode, Konstantia Asteriadou, Phillip T. Robbins, and Peter J. Fryer

Abstract: Fouling of food process plant surfaces and the subsequent cleaning needed is a significant industrial problem,
and as the cost of water and chemical disposal increases, the problem is becoming more significant. Current literature
on water-based cleaning is reviewed here according to the classification of 3 types of cleaning problems. By doing this,
it is hoped that new knowledge can be highlighted applicable to improving industrial cleaning. (i) For type 1 deposits
(that can be cleaned with water alone)—Cleaning time appears related to Reynolds number and surface shear stress. An
increase in Reynolds number seems to decrease cleaning time. Cleaning temperatures greater than 50 ◦ C do not appear
beneficial. (ii) For type 2 deposits (biofilms)—Removal behavior of biofilms seems to be dependent on the microbial
aging time on the surface. Keeping a material hydrated on a surface enables easier removal of it with water. a. Water
rinsing: Temperature and wall shear stress have varied effects on removal. b. Chemical rinsing: Flow and temperature were
seen to have the biggest effect at the start of cleaning, but contact time was more important as cleaning progressed at a
given sodium hydroxide solution flow and temperature. (iii) For type 3 deposits (that require a cleaning chemical)—For
specifically, protein-based systems excessive chemical forms a deposit difficult to remove. Increasing wall shear stress and
temperature was most beneficial to cleaning rather than concentration. The action of temperature can reduce the use
of a chemical for type 2 and type 3 soils. The findings suggest that the right combination of flow characteristics at a
given temperature and concentration is crucial to achieving fast cleaning in all cases. There are a number of cleaning
monitoring methods at various stages of commercialization that may be capable of monitoring bulk cleaning and cleaning
at the surface. To optimize cleaning will require integration of measurement methods into the cleaning process.

Contents 3.3.3 Chemical effects on the cleaning of


type 2 deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 3.3.4 Chemical effects on the cleaning
2. Fouling studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 of type 3 deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.1 Adhesion of microorganisms to surfaces . . . . . . . . . . . . . . . . 4 4. Novel cleaning approaches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.3 Preventing fouling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 4.1 Increasing boundary layer disruption. . . . . . . . . . . . . . . . . .16
2.3.1 Process surface modification . . . . . . . . . . . . . . . . . . . . . . 5 4.2 Alternative cleaners to reduce environmental impact . . . 16
2.3.2 Process alterations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6 4.3. Other studies related to cleaning behavior . . . . . . . . . . . . 17
3. Cleaning. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6 4.3.1 Deposit shear . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.1 Product recovery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 4.3.2 Deposit deformation and strength . . . . . . . . . . . . . . . . 17
3.2 The effect of CIP parameters on type 1 removal . . . . . . . 10 5. Measuring cleaning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
3.2.1 Flow and wall shear stress . . . . . . . . . . . . . . . . . . . . . . . . 11 5.1 Online bulk measurements . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.2.2 Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 5.2 Online surface measurements . . . . . . . . . . . . . . . . . . . . . . . . 19
3.2.3 Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 5.3 Measuring microbial cleanliness . . . . . . . . . . . . . . . . . . . . . . 19
3.3 The effect of CIP parameters on type 2 and type 6. Summary and conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
deposit removal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.3.1 Membrane cleaning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.3.2 Water-rinsing of hard surfaces . . . . . . . . . . . . . . . . . . . . 14
Introduction
The success of a branded fast moving consumer goods (FMCGs)
business depends fundamentally on product quality and safety con-
MS 20120483 Submitted 3/29/2012, Accepted 10/14/2012. Authors are with formance at a required level. Poor cleaning or hygiene confor-
School of Chemical Engineering, Univ. of Birmingham, Edgbaston, Birmingham, B15 mance can be a result of fouling layers building up in a plant
2TT, U.K. Direct inquiries to author Fryer (E-mail: p.j.fryer@bham.ac.uk). or other problems. Figure 1 illustrates a typical route by which
a large FMCG manufacturer will ensure a hygienic plant. There


C 2013 Institute of Food Technologists®

doi: 10.1111/1541-4337.12000 Vol. 12, 2013 r Comprehensive Reviews in Food Science and Food Safety 121
Critical review in fouling and cleaning . . .

Scienfic invesgaon Define CIP standard


• Invesgate design • Best process design
• Invesgate process parameters • Best process parameters

LACK OF FUNDAMENTALS
e.g. The effect of CIP
parameters on different fouling

Input from pracce


condions Establish standard at all sites
By site educaon, training and
Input from science

Input from pracce


empowerment and recording site
NON - CIP condions
CONFORMANCE
e.g. Poor
maintenance,
control or training
Is hygiene
Why
NO being YES
not?
achieved?

Figure 1–Flow diagram illustrating the route taken by industry to ensure plant hygiene (Heineken personal communication 2012).

should be fundamental research and development obtained from identify the best way to clean a processing plant from experiments
both practice and science that are integrated and applied in-plant of different plants. The direct selection of cleaning protocols is
to provide the optimum cleaning protocol. In food and bever- not always possible. In practice, cleaning protocols can only be
age manufacturing operations, cleaning-in-place (CIP) is used to developed semiempirically in industry. In most cases, CIP cannot
remove residual product, fouling, and microbes that remain in be optimized in situ because of the risk posed of compromising
the process line from production. The act of cleaning therefore existing cleanliness.
maintains product quality, safety, and production efficiency. Dur- Fryer and Asteriadou (2009) suggest a classification of cleaning
ing CIP, water and/or chemical solution is circulated around plant problems in terms of cleaning cost and soil complexity. A diagram-
process equipment. With large-scale manufacturers, the process is matic representation of this relationship is presented in Figure 2.
generally fully automated. A typical CIP philosophy in industry is This classification enables the nature of a foulant to be related to
that of Scottish & Newcastle Breweries (2008): the type of cleaning employed, and therefore, the cost. This clas-
sification also indicates the environmental impact of the type of
“ensure all production, processing, and packaging plant is cleaning employed; complex soils require chemical and thermal
cleaned by a standard regime and to a schedule which ensures cleaning that lead to a high cleaning cost and high environmen-
cleanliness and microbiological integrity at all times; with min- tal impact. Three deposit types were chosen to represent a broad
imum cost, energy, and delay to production in a manner which range of cleaning problems seen in food, beverage, and personal
ensures human, plant, product, and environmental safety.” care products manufacturing:

A significant body of cleaning knowledge exists within individ- (i) Type 1: Viscoelastic or viscoplastic fluids such as yogurt and
ual manufacturers, equipment suppliers, and chemical companies; toothpaste that can be rinsed from a process surface with
however, the determined cleaning regimens have often been kept water.
confidential and plant-specific. This has resulted in independent (ii) Type 2: Microbial and gel-like films such as biofilms and
development of cleaning operations. Organizations such as the polymers removed in part by water and in part by chemical.
European Hygienic Engineering and Design Group (EHEDG) (iii) Type 3: Solid-like cohesive foulants formed during ther-
have produced extensive guidelines on the types of surface and mal processing such as milk pasteurization and brewery
equipment that are easy to clean, such as detailed in the EHEDG wort evaporation. These operations mostly require chemical
Yearbook (2007). removal.
CIP tends to follow a similar series of steps for a prescribed Yang and others (2008) also classified cleaning optimization
time and at a prescribed flow rate, temperature, and chemical methods, here into 2 types of investigation:
concentration known to give a repeatable level of cleanliness. It
is not yet possible to predict before an operation how a given (i) Engineering investigations: reducing energy, time, and cost in
piece of equipment could foul and be cleaned. It is difficult to established cleaning operations.

122 Comprehensive Reviews in Food Science and Food Safety r Vol. 12, 2013 
C 2013 Institute of Food Technologists®
Critical review in fouling and cleaning . . .

Figure 2–Cleaning map; a classification of cleaning problems based on soil type and cleaning chemical use (from Fryer and Asteriadou 2009).

(ii) Scientific investigations: achieving cleanliness or a clean- r Growth of biofilm—often after the formation of a “condi-
ing time as a function of influencing factors; for ex- tioning layer’’ of protein onto the surface.
ample, wall shear stress, temperature, surface type, and r Accumulation of material in stagnant or low-flow areas of
finish. equipment.
r Loss of membrane activity.
The aim of this review was to provide an overview of current
knowledge on cleaning solutions classified by soil cleaning type. Fouling is a costly problem in the food, beverage, and other
This novel classification is hoped to highlight new CIP optimiza- industries, which is often unavoidable due to the heat treat-
tion opportunities for industry and any future research in the ment that often has to be given to products to develop cer-
field. Current knowledge of fouling prevention and novel clean- tain colors and flavors and ensure safety. By definition, foods
ing methods are also discussed here. are sources of nutrients favorable not only to people but also
to microbes that stick to process surfaces—so microbial ad-
Fouling Studies hesion to surfaces and subsequent growth are important phe-
Fouling is defined as the unwanted buildup of material on a nomena. The economic penalties of fouling in heat exchang-
surface. The fouling process generally involves a number of steps ers were discussed by Müller-Steinhagen (2000) and can be
(Epstein 1983): summarized:

(i) surface conditioning, (i) Capital expenditure, due to:


(ii) mass transfer of species to the surface, (a) Excess heat transfer surface area compensating for the occur-
(iii) surface deposition, rence of fouling. This has been estimated as an average
(iv) deposit aging, and of 30% additional capital cost.
(v) possible removal. (b) Higher transport and installation costs for bigger and heavier
equipment.
There is also a classification of fouling mechanisms demonstrated
(c) Cleaning systems, including their installation and mainte-
by Bott (1990) detailed in Table 1. Fouling problems that have
nance costs.
been reported in the food and beverage industries include (but
this is by no means a complete list): (i) Fuel cost—If extra energy (such as steam) is required to
keep the fouled heat exchanger operating for the required
r Protein and mineral deposition in heat exchangers. performance.
r Ice buildup in freezers. (ii) Maintenance cost—Of the heat exchanger, cleaning sys-
r Scale buildup in cooling water systems. tem, and any ancillary equipment in the process (and clean-
r Fat burn-on in ovens. ing) loop, for example, chemical tank level probes, flow
r Product solidification. meters, interface probes, and boilers.


C 2013 Institute of Food Technologists® Vol. 12, 2013 r Comprehensive Reviews in Food Science and Food Safety 123
Critical review in fouling and cleaning . . .

Table 1–Fouling mechanisms: adapted from Bott (1990) and Sharma and others (1982).

Fouling mechanism Underlying process


Crystallization Formation of crystals on the surface formed from solutions of dissolved substances when the solubility limit is changed.
Cooled surfaces are subject to fouling from normally soluble salts, fats, and waxes. Inversely soluble salts, such as calcium
phosphate deposits on heated surfaces. Where the fluid or components of the fluid solidify onto the surface, this is called
solidification fouling (Sharma and others 1982).
Particulate deposition Small suspended particles such as clay, silt, or iron oxide deposit onto heat transfer surfaces. Where settling by gravity is the
determining factor, this is then called sedimentation fouling.
Biological growth (biofouling) The deposition and growth of organic films consisting of microorganisms and their products, called biofilm.
Chemical reaction at Reaction of some part of the flow to generate insoluble material. The deposit formed on the surface (particularly heat
fluid/surface interface transfer surfaces) has a different composition to the process fluid (for example, in petroleum refining, polymer production,
and dairy plants).
Corrosion The material of the heat transfer surface is involved in reactions with components of the fluid to form corrosion products on
the surface, a specific type of chemical reaction fouling.
Freezing Deposit formed from a frozen layer of the process fluid, for example, ice from water or solid fats from a food fluid.

(iii) Cost due to production loss—Cost of continuous pro- organisms usually follows the formation of a conditioning layer
duction (without shut-down for cleaning or maintenance) of protein (Lorite and others 2011) that makes subsequent adhe-
as compared to the actual production cost. sion and biofilm formation easier. The sequence of events that
occur during film formation is discussed by Busscher and others
Accurate measurement of the effects of fouling and the efficiency
(2010) and Chen and others (2010) who show the kinetics of film
is critical. Changes in heat transfer efficiency have widely been
formation.
recorded. Most common is following the change in heat transfer
Other researchers have studied yeast adhesion and proliferation
during fouling by including a fouling resistance, Rf , in the equation
on processing surfaces, critical in brewing operations. Reynolds
relating the initial clean heat transfer coefficient, (U0 ), to that at
and Fink (2001) proved that Baker’s yeast can initiate biofilm for-
time t, (U):
mation on plastic when in a low-glucose environment. Mozes and
1 1 others (1987) found that yeast could attach and form a dense layer
= + Rf (1)
U U0 of cells on stainless steel and aluminum at pH 3 and pH 5 and 6.
The authors also determined that a dense layer of yeast cells would
And the extent of fouling may be expressed by a Biot number attach to glass and plastics if the negative charge was reduced by
(Bi), which accounts for deposit thickness (x) and thermal con- treatment with ferric ions. The system pH will determine the
ductivity (λ): Bi = Rf .U0 , where Rf = x/λ for the deposit. Deposit surface charge of both the substrate and the adhering species. The
resistance during cleaning can be described as the reverse process isoelectric point, the pH where the material carries no charge, will
to (1) (Tuladhar 2001) as also vary with surface and organism. Yeast has also been found by
1 1 other authors to readily attach to stainless steel, plastics, elastomers
Rd = − (2) (Guillemot and others 2006), and glass (Mercier-Bonin and oth-
Uc Ut
ers 2004), all of which are used extensively in FMCG industries.
where Ut is the heat transfer coefficient at time t and Uc the heat The effect of cleaning parameters on yeast removal from process
transfer coefficient of the final clean system, so the rate of change surfaces is discussed in later sections.
of this is a measure of cleaning. Product contact surface finishes with a roughness (Ra ) value
The rate and extent of fouling and cleaning is often classified of up to 0.8 μm are recommended (Lelieveld and others 2005),
in terms of fluid flow, either in terms of the Reynolds number which is often called 2B finish of stainless steel. Surface rough-
(Re = ρvd/μ, where ρ and μ are the density of viscosity of a ness exists in 2 principal planes, one perpendicular to the surface
fluid flowing at mean velocity v through a system of characteristic described as height deviation and one in the plane of the sur-
length d, such as pipe diameter) or the surface shear stress. In face described by spatial parameters. The effect of average surface
this paper, many correlations in terms of Reynolds number are roughness height, Ra , and surface topography on microbial re-
discussed—to convert to velocity requires knowledge of density tention has been investigated most thoroughly. Hilbert and others
and viscosity of the fluid, which is simple for water but may be (2003) investigated the effect of stainless steel roughness (Ra 0.9 to
more complex for cleaning solutions. 0.01 μm) on retaining various microbes. The surfaces also had a
conditioning layer. The retention of microbes (measured by indi-
Adhesion of microorganisms to surfaces rect conductometry) on the conditioned surfaces was similar over
The principal factors responsible for adhesion between surface the range of Ra tested.
and foulant include: (i) van der Waals forces, (ii) electrostatic forces, Cluett (2001) investigated the effect of stainless steel surface
and (iii) contact area effects; the larger the area, the greater the finish on the fouling and cleaning of a beer fermenter. Surface
total attractive force (Bott 1995). Microbes have a natural affinity finishes investigated included 2B milled stainless steel and me-
to surfaces. Numerous authors have reported the adhesion of bac- chanically polished 120 grit, 240 grit, and electropolished (EP)
teria to processing surfaces (for example, Geesey and others 1996; stainless steel. The top surface of the fermenter was half EP, half
Bénézech 2001; Zhao and others 2007). If left to proliferate, in- 240 grit, and the cone was EP. The cylinder of the vessel had
dividual microbes can grow into biofilms (adhesive and cohesive all finishes, one quarter of the vessel from top to bottom repre-
communities of microbes) that become difficult to remove from sented by each surface finish. After lager beer fermentation lasting
a surface (Jefferson 2004). Garrett and others (2008) summarize 12 d, Cluett (2001) found that all surfaces fouled similarly and
the occurrence of biofilms in industry, fouling mechanisms and the level of deposition was heavy. He also found that all the sur-
methods of observing and probing structures. The adhesion of faces cleaned similarly using a similar CIP regime with a spray ball

124 Comprehensive Reviews in Food Science and Food Safety r Vol. 12, 2013 
C 2013 Institute of Food Technologists®
Critical review in fouling and cleaning . . .

yeast adhesion. Overall surface energy was 50 to 60 and 25 mJ/m


(prerinse, caustic, water, acid, water, and sanitizer). However, the
number of viable microbes was found to decrease in the cone at for cell adhesion on plastic and glass, respectively. Coating both
the bottom of the vessel. surfaces changed the free energy of the system resulting in a de-
Gallardo-Moreno and others (2004) investigated the effect of crease to 35 to 40 mJ/m for plastic and an increase to 30 to 40
surface roughness by comparing yeast adhesion to glass (Ra 0.8 mJ/m for glass. Yeast adhesion was significantly reduced on the
μm and hydrophilic) and silicone rubber (SR) (Ra 0.61 μm plastic surface coated in the peptide and increased on the glass
and hydrophobic). The authors found larger adhesion rates for surface. Changes in surface roughness and hydrophobicity due to
SR, and at 37 ◦ C rather than 22 ◦ C. Whitehead and others the coating will also have contributed to adhesion.
(2006) investigated Pseudomonas aeruginosa (rods of 1 μm width and Quain and Storgårds (2009) mentioned the testing of “func-
3 μm length) and Staphylococcus aureus (1 μm sphere) retention on tional materials” in the lab and in brewery dispense lines such
a titanium dioxide surface: smooth with defined surface features as hydrophobic fluoropolymer coatings, photocatalytic titanium
(pits) of 0.5 μm. S. aureus cells were removed more easily from the
dioxide coatings, and the inclusion of antimicrobial silver ions
smooth surface, whereas P. aeruginosa cells were removed more eas-(0.042%) in stainless steel. The latter was shown to reduce the
ily from the defects. Whitehead and Verran (2006) also reviewed number of adhering bacteria by 99% compared to normal stainless
the effect of Ra and topography on microbial retention. Research steel. However, the effect decreased with time.
suggests that surfaces with a Ra value close to the cell size see The influence of surface energy on adhesion is well known in
increased retention on the surface. For example, yeasts were foundmarine and medical biofouling and is characterized by the “Baier
to require larger defects (5 μm) for retention and smaller daughter
curve” (Baier 1980). This curve demonstrates the weakest adhesive
cells were retained in smaller defects (2 μm). Rod-shaped cells strength of bacteria to be at surface energies of around 25 mN/m.
seemed to orient themselves in grains and grooves of similar size. Equations defining possible minimum adhesion energies be-
tween a deposit and the surface have been developed. The follow-
Preventing fouling ing equation has been derived:
If fouling were not to occur, there would be little need for    
cleaning. Broadly speaking, 2 methods for preventing fouling have 1
γ SLW = γ DLW + γ FLW (3)
been approached in the literature: 2
(i) Functional surfaces—For example, smooth surfaces with spe-
where γ SLW , γ DLW , and γ FLW are the Lifshitz–van der Waals (LW)
cific finish, topography, hydrophobicity, or surface charge.
surface free energy of the surface, deposit, and fluid, respectively,
“Nonstick surfaces” are designed to have a specific surface
and which can be quantified from contact angle measurements
energy to minimize fouling.
(Zhao and others 2004). Liu and others (2006) studied the inter-
(ii) Processing alterations—For example, changing product flow
actions of 316 L stainless steel with baked and unbaked tomato
characteristics, holding times, transient times, and other pro-
deposit: a minimum removal energy range of 20 to 25 mN/m was
cess parameters designed to minimize fouling.
found in both cases. Either side of this surface energy range, the
Process surface modification. A hygienic surface needs to be adhesive strength of the deposit on the surface increased. Zhao
smooth, easy to clean, able to resist wear, and retain its hygienic and others (2005a) found that stainless steel surfaces coated with
qualities. Stainless steel is the most common food contact material Ag-PTFE reduced Escherichia coli attachment by 94% to 98%, com-
used in the industry, being stable at a variety of temperatures, inert, pared with silver coating, stainless steel, or titanium surfaces. A sur-
relatively resistant to corrosion, and it may be treated mechanically face with energy of 24.5 mN/m roughly matching the theoretical
or electrolytically to obtain a range of finishes (Akhtar and others minimum adhesion energy of E. coli, 28.3 mN/m, was achieved.
2010). The wettability of a surface is dependent on its surface Composite coatings using nickel, phosphorus, copper, and PTFE
energy. A surface with a high surface energy is hydrophilic and a were also used by Zhao and others (2005b) and Zhao and Liu
drop of cleaning fluid will spread over the surface. A low-energy (2006) to create surfaces with specific energies shown to reduce
surface is hydrophobic and a drop of water will not spread. Water biofouling. A major EU project (“MODSTEEL”) developed and
partially wets glass and acrylic and does not wet Teflon (PTFE) studied a wide range of surfaces and how they might reduce foul-
surfaces—but surfactants are often added to commercial cleaning ing from milk (see Santos and others 2004 and Rosmaninho and
agents to improve wetting. Wetting is determined by the nature of others 2007).
both the liquid and the solid substrate. The cleanability (and disin- Work by Pereni and others (2006) confirmed the effect of
fectability) of stainless steel has been compared with those of other surface free energy in minimizing P. aeruginosa adhesion to a
materials, and is comparable to glass when cleaning microbes, and range of coatings including silicone, polished and nonpolished
significantly better than polymers, aluminum, or copper (Akhtar stainless steel, PFA (perfluoroalkoxy polymer) and PTFE nickel,
and others 2010). phosphorus, and aluminum composite coatings. The total surface
Microbes are known to readily attach to SR. Everaert and others free energy was in the range 17.2 to 48.3 mN/m, as shown in
(1998) absorbed long fluorocarbon chains (Ar-SR-C8 F17 ) to SR Figure 3(A). Minimum retention of bacteria was found at 20 to
used in prosthetics in an attempt to reduce the number of adhering 27 mN/m. Silicone had a surface free energy of around 20 mN/m
microbes. They found that the initial adhesion rate of Streptococ- and the lowest colony forming units (CFUs) count. Surface free
cus bacteria to the treated rubber was significantly reduced, from energy has been shown to be the parameter dominating E. coli
around 2500 to 900 cm−2 s−1 , without a conditioning film of adhesion over a range of metal–polymer coatings, and a minimum
saliva and 400 cm−2 s−1 with a conditioning film of saliva. The adhesion energy of 25 mN/m was found (Zhao and others 2007),
adhesion rate of Candida species to treated rubber was also reduced as shown in Figure 3(B).
compared to untreated rubber. Parbhu and others (2006) used a transient treatment to modify
Dhadwar and others (2003) investigated the effect of oligopep- a stainless steel surface. The treatment was present during the pro-
tide treatment of glass (hydrophilic) and plastic (hydrophobic) on cessing cycle and removed at high pH during alkaline cleaning.


C 2013 Institute of Food Technologists® Vol. 12, 2013 r Comprehensive Reviews in Food Science and Food Safety 125
Critical review in fouling and cleaning . . .

cell; a potential survival strategy. Treatment of water systems with


silver nanoparticles could prevent significant biofilm buildup.
Tse and others (2003) found that in a 2-phase (liquid-vapor)
wort boiling system, the wall temperature did not significantly
affect the rate of fouling. Under conditions where vapor was con-
densed at lower flow velocities (0.07 and 0.14 m/s), the initial
fouling phase was more rapid at the higher flow velocity. The
authors found that the initial fouling rate was halved as the flow
velocity was doubled. These findings suggest that circulating fluid
at a fast flow rate would reduce fouling. The authors also found
that at the lowest flow rate, 0.07 m/s, and highest temperature,
170 ◦ C, the foulant appeared the most severe. The fouling also
had different makeup depending on its position in the column.
At the top of the column, the deposit was light in color, smooth,
and patchy; at the bottom, the deposit was dark brown and mul-
tilayered. The authors suggested 2 fouling mechanisms: chemical
reaction of species in the wort forming polymers and crystallization
of species from the wort due to evaporation at bubble nucleation
sites in nucleate boiling regions. Liu and others (2004) compared
fouling of 2-phase flow (liquid-vapor) and 3-phase flow (liquid–
vapor–solid) during the evaporation of Gengnian’an extract. The
solid phase was added as inert solid particles. The 2-phase flow sys-
tem generated fouling in 15 h, whereas the 3-phase flow system
generated fouling after 60 h.
Modifying the process by using electric fields has also been dis-
cussed. Ohmic heating occurs when an electric current is passed
directly through milk to heat it, rather than it being heated by
surface heat transfer. The process results in lower surface temper-
atures and less fouling initially. However, fouling in the bulk is
easily transferred to the surface, resulting in fouling (Bansal and
Figure 3–(A) P. aeruginosa AK1 retention on investigated surfaces
compared with the total surface free energy (adapted from Pereni and
Chen 2006). Kim and others (2011) demonstrated that an elec-
others 2006). (B) Effect of surface free energy on E. coli adhesion tric field could be used to control membrane fouling with E. coli.
(adapted from Zhao and others 2007). E. coli cell suspensions were treated by an electric field prior to
filtration. The flux of the suspension was maintained through-
out the filtration period due to larger fouling particles reducing
cake resistance. Cell death also increased with increasing electric
The treatment was shown to reduce the interaction potential be- field strength from 5 to 20 kV/cm. Flux of the untreated E. coli
tween stainless steel and phosphate anions resulting in significant suspension decreased abruptly after the onset of filtration.
reductions in fouling rates. Xiaokai and others (2005) investigated the effect of electro-
Akhtar and others (2010) compared adhesion of a range of magnetic treatment of water to minimize scale formation in the
food and personal care foulants to different surfaces. Particle tips tubes of a plate heat exchanger (PHE). The technology is termed
of different materials were attached to an atomic force micro- electromagnetic antifouling (EAF). The treatment was shown to
scope (AFM) cantilever to study the detachment from toothpaste aggregate particles in the flow that led to reduced precipitation at
and some confectionery components: Turkish delight, caramel, the wall.
and sweetened condensed milk (SCM). The study did reveal sig-
nificantly different detachment forces for the same deposit from Cleaning
different surface types (see Figure 4). Caramel and SCM seemed No economically viable fouling prevention method is yet to
to be more difficult to detach from glass than stainless steel. It was be demonstrated in industry. Should one of the modified sur-
possible to relate data from the AFM to measurements taken on face methods prove economic, then the problem will be greatly
a millimeter scale using micromanipulation probes (Liu and oth- reduced. Understanding the cleanability of surfaces requires com-
ers 2002, 2006, 2007). Akhtar and others (2012) describe further bining understanding of surface chemistry and engineering, the
research using AFM to study food adhesion to process surfaces. deposit and the cleaning fluid (for a recent review of clean-
Process alterations. Dror-Ehre and others (2010) tested the ef- ability, see Detry and others 2010). Further research discussed
fect of biofilm development of P. aeruginosa when pretreated in an here considers the findings of studies relevant to optimizing
aqueous solution of molecularly capped silver nanoparticles (MC- cleaning.
NPs). Under specific conditions, cells and surfaces incubated for One significant issue is determination of the correct cleaning
39 h at 37 ◦ C, Ag-MCNPs retarded biofilm formation even when time. A deposit that has aged on a surface is more difficult to
a high percentage of planktonic P. aeruginosa cells survived pretreat- remove than fresh material on a surface, so cleaning is encouraged
ment with Ag-MCNPs. At the various incubation times, a stable, after production. Aging of a particular soil type could make a
low value of biomass was formed that could be easily removed. deposit harder to remove from a process surface. For example, a
The authors found, from micrographs of pretreated cells, that the type 1 soil could become a type 2 soil over time and heating may
intracellular material was pushed toward the peripheral parts of the result in a type 3 soil. Goode and others (2010) found that in

126 Comprehensive Reviews in Food Science and Food Safety r Vol. 12, 2013 
C 2013 Institute of Food Technologists®
Critical review in fouling and cleaning . . .

1 Stainless steel
Caramel SCM Turkish Toothpaste PTFE
delight
Glass
0.1

F/R(N/m)
0.01

0.001

0.0001

Figure 4–Force of attraction between stainless steel, PTFE (fluorinated low energy surface), and glass particles and different food materials obtained
using AFM (from Akhtar and others 2010). F/R is the force/probe radius with units of N/m.

beer fermentation vessels, there were 2 distinct deposit types to be and Asteriadou 2009), is a useful cleaning problem classification
cleaned, classified as type A and type B foulants: tool and forms the basis for the structure of this review. Examples
(i) Type A—Formed during fermentation above the beer level of each deposit type include:
at the top of the vessel, (i) Type 1: toothpaste, tomato paste, yogurt, shampoo, beer,
(ii) Type B—Residual yeast attached to the vessel wall and cone wine, milk, and yeast.
below the beer level during emptying. (ii) Type 2: microbes and microbial films of bacteria, spores, and
yeast species.
Type B fouling is shown by Salo and others (2008), as seen in (iii) Type 3: milk, whey protein concentrate (WPC), cooked
Figure 5(A), while an example of type A foulant viewed from a SCM, starch, boiled wort, and egg albumin.
fermenter man way door at the top of the vessel is given in Figure
5(B). Type B fouling has a shorter aging time than type A fouling. Some of the research that has considered the influence of clean-
As such, type B foulant can be removed by the falling film in a ing parameters in flowing systems on the removal behavior of
tank, whereas type A foulant may require a larger impact force deposits is listed in Table 2 to 4. Table 2 details type 1 deposit
for removal or a combination of water and chemical rinses for removal studies, Table 3 details type 2 deposit removal studies,
complete removal (Goode and others 2010). Similarly, Liu and and Table 4 details type 3 deposit removal studies. The cleaned
others (2002) found that the force required to remove a tomato geometry, effect of CIP parameters, and the method of deter-
deposit from a surface increased with time until after about 200 mining cleaning effectiveness are listed in each Table. The ef-
min of heating it remained constant. fect of flow has been studied both in terms of the Reynolds
Automated CIP has been widely applied in dairies, food pro- number (Re) and the surface shear stress. Both may provide
cessing, brewing, and wine processing for the last 50 y to return further insight into the effect of removal behavior on flow
the plant to a clean state (Stewart and Seiberling 1996). Dur- velocity.
ing CIP, water and chemicals are circulated around the plant for Milk processing is a large industry and fouling is a significant
a prescribed duration (Tamine 2008). The CIP factors found to problem, as both protein aggregates and minerals are deposited;
determine cleaning can be described by Sinner’s circle, a circle Burton (1967) classified the proteinaceous deposit seen in pasteur-
of the cleaning parameters: mechanical action, chemical action, izers as type A and the mineral deposit seen at UHT temperatures
time, and temperature (Lelieveld and others 2005). Cleaning can as type B. Reviews of dairy fouling research are presented by
also be dependent on geometry. In a pipe, the contribution of the Changani and others (1997) and Bansal and Chen (2006). Pro-
cleaning factors is equal. In a pipe dead leg, time determines clean- teins have been identified as a major source of fouling deposits.
ing (Lelieveld and others 2005). A number of attempts have been Fickak and others (2011) found that increasing the protein con-
made to try to incorporate computational models into the design centration of whey protein increased the amount of fouling on a
process, as shown by Asteriadou and others (2006) and Jensen pilot-scale heat exchanger. Holding of milk before heating sec-
and Friis (2005). This approach will become more important as tions has been shown to aggregate β-lactoglobulin in the holding
understanding of the processes in cleaning increases. sections rather than the heating sections (de Jong and van der Lin-
Rheological characterization of materials enables their classifi- den 1992). Christian and others (2002) found that increasing the
cation. Materials within a similar class may have similar cleaning mineral content of whey protein decreased the extent of fouling
behavior, according to the classification by Fryer and Asteriadou on a PHE.
(2009). Vinogradov and others (2004) characterized the rheology WPC is often used in research studies to represent a milk fouling
of a dental plaque biofilm. Biofilm rheology has been viscoelas- deposit, because it is easier to handle and store than milk, and the
tic, temperature-dependent and/or time-dependent (Rao 1999). fouling composition is thus easier to control and replicate. Robbins
Characklis (1980) compared the elastic and viscous modules ob- and others (1999) compared the cleaning of milk and WPC from
tained for a biofilm and a cross-linked protein gel, fibrinogen. The a PHE. They found that in the pasteurization and UHT sections
elastic modulus was the same order of magnitude for the protein of the PHE, both materials fouled heavily. However, in the inter-
gel and the biofilm. The cleaning map, presented in Figure 2 (Fryer mediate section, WPC also fouled excessively, whereas milk did


C 2013 Institute of Food Technologists® Vol. 12, 2013 r Comprehensive Reviews in Food Science and Food Safety 127
Critical review in fouling and cleaning . . .

A B Gasket

T
Type A fouling on the
v
vessel he gasket
and th
Wall fouling (s d been
samples had
Contactt agar s
scraped from the
vessel)
v

Fermente
er
cone

Figure 5–(A) 80 L stainless steel tank (0.8 m × 0.4 mm) with residual yeast fouling attached to the wall and the cone. The wall was also sampled by
contact agar (adapted from Salo and others 2008). (B) Type A deposit seen at the top of a fermenter around the man way door and the gasket (Goode
2012).

Table 2–Some CIP studies of type 1 deposit.

Effect of Effect of Cleaning


Deposit Geometry flow or τ w temperature Effect Re determinant Reference
Toothpaste 1 m long, 2 OD, 316 L Increase flow Increase Increase Re (4000 Turbidity reaches Cole and others
ss pipe (horizontal) velocity (1 to 3 temperature to 250000) 4 ppm. (2010)
m/s, decrease (from 20 ◦ C), decreases
cleaning time. decrease cleaning time.
cleaning time
(to a point
approximately
40 ◦ C).
Shampoo 316 ss plate (350 mm (0.14 to 0.47 m/s) (31 to 51 ◦ C), — Visual MSS and Pereira and others
long, 30 mm ID, higher flow removal of spectrophotom- (2009)
18.3 mm ED) velocity, more shampoo layers etry
(vertical flow cell) efficient faster at higher
removal at the temperatures as
start of cleaning
cleaning. proceeds.
Mustard glass T-piece (variable Increase flow — Above a certain Visual Jensen and others
depth T, 4 and 6 velocity (1 to Reynolds (2007)
cm) 1.88 m/s) number, the
increase recirculation
removal rate. zone length
becomes
constant.
Yeast cells glass, polypropylene, Increase τ w , — — Visual Guillemot and
rehydrated and polystyrene decrease others (2006)
(aged 1 h at surfaces (210 × 90 number of cells.
ambient) mm long) in
horizontal flow cell
(i) linearly for
plastics
(ii) as a curve for
glass
Tomato paste 316 L ss coupons (0.7, 1.5, 2.3 (30, 50, and 70 (850 to 4800 Re) Visual, image Christian (2004)
(circular: 26 mm D) L/min) increase ◦ C) increase increase Re, analysis, and
horizontal flow cell flow rate, the temperature, decrease MHFS
effect of decrease the cleaning time
temperature on time to remove
cleaning time deposit
decreases
OD = outer diameter, ID = inner diameter, ED = equivalent diameter, ss = stainless steel, MSS = mechatronic surface sensor, MHFS = microfoil heat flux sensor.

not. Compositional analysis revealed protein fouling from both others (2006) found that yeast cells could be wholly removed from
materials in the pasteurizer section. Increasing to UHT tem- glass using water but that yeast cells had strong adhesion to stainless
peratures revealed milk fouling to become more mineral-based, steel. The wall shear stress required to remove 50% of the attached
whereas the WPC fouling remained predominantly protein-based, cells from stainless steel, denoted as τ w50% , was 30 Pa, while for
suggesting comparison of milk fouling and WPC fouling is not plastics τ w50% ranged from 1 to 2 Pa.
wise at UHT temperatures. The effect of CIP parameters on the removal of different de-
Yeast can exhibit type 1 (if in contact with glass) and type 2 (if posit types is discussed in the following sections. Even though
in contact with stainless steel) cleaning behavior. Guillemot and there is clear evidence that different deposit types are removed

128 Comprehensive Reviews in Food Science and Food Safety r Vol. 12, 2013 
C 2013 Institute of Food Technologists®
Critical review in fouling and cleaning . . .

Table 3–Selected CIP studies of type 2 deposit.

Effect of Effect of Effect of Cleaning


Deposit Geometry flow or τ w temperature chemical/pH determinant Reference
Yeast slurry (aged 316 ss coupons Increase in flow Increase 1% NaOH Visual, image Goode and others
at 30 ◦ C, 5 d) (square: 30 × velocity (0.26 to temperature, analysis, and (2010)
30 mm long) in 0.5 m/s) decrease MHFS
horizontal flow decrease cleaning time.
cell cleaning time at
50 and 70 ◦ C.
Limited effect
beyond 0.4 m/s
at 20 and 30 ◦ C.
B. cereus spores 316 L ss pipe (20 — — NaOH 0.5% Agar overlay Le Gentil and
cm long, 2.37 (w/w) at 60 ◦ C, technique using others (2010)
cm ID) and 2 2200 L/h, up to TTC (spores
way valve (entry 30 min. The % appear red)
to exist 18 cm residual spores
long, 3.5 cm ID) decreased as
cleaning time
increased.
Yeast cells 316 L ss, (210 × Increase τ w , — — Visual Guillemot and
rehydrated 90 mm long) in decrease others (2006)
(aged 1 h at horizontal flow number of cells
ambient) cell barely for
stainless steel
(10%).
B. cereus spores (in 304 L ss pipes (15 Increase τ w Rinsing at 60 ◦ C 0.5% w/w of Agar overlay Lelièvre and
milk) × 10−2 m long, (17.45 to 68.95 revealed less NaOH at 60 ◦ C technique using others (2002)
2.3 × 10−2 m Pa, that is, 1.61 spores TTC (spores
ID) (horizontal) to 3.29 m/s) compared to 20 appear red)
decrease ◦ C at the same
number of soaking times.
spores (after 5
min). Contact
time was more
important in
reducing spores.
B. cereus spores (in Progressive-cavity — — Prerinse 0.5 m/s Agar overlay Bénézech and
custard) pump (with (6 min); 0.2% technique using others (2002)
axial or NaOH at 1.5 TTC (spores
tangential exit m/s, 60 ◦ C (10 appear red)
pipe). min);
Tangential was intermediate
best. In the rinse 0.5 m/s (6
axial setup, the min); 0.2%
number of CFU HNO3 at 1.5
was greater m/s, 60 ◦ C (10
than 10 min); final rinse
CFU/cm in the 0.5 m/s (6 min).
pump body and
gaskets.
OD = outer diameter, ID = inner diameter, ED = equivalent diameter, ss = stainless steel, MSS = mechatronic surface sensor, MHFS = microfoil heat flux sensor.

from surfaces differently, the approach to cleaning is typically the Product recovery
same: At the end of a process, there can be a significant amount of
material left in pipes and tanks. This product may be saleable, in
(i) Prerinse (or product recovery stage); to remove loosely which case it should be recovered, or it may be considered waste. In
bound soil and product. both cases, the bulk of this material should be removed (generally
(ii) Detergent phase (alkali or acid); to remove the fouling layers. in the first rinse phase) prior to the “cleaning” phase. Type 1
(iii) Intermediate rinse; to remove chemical. deposits will generally be saleable for products like toothpaste or
(iv) Sanitization/disinfection step (chemical and/or thermal); to shampoo, and recovery should be maximized. Type 2 and type 3
kill viable microbes and restore the hygienic condition of deposits will generally be formed as thin layers at the wall of a
the system. different composition to product. The layers need to be removed,
(v) Final water rinse; so product can be reintroduced to the to return the plant to a clean state, and tend not be recovered at
system. the end of a process.
Palabiyik and others (2012) investigated the effect of the prod-
Although disinfection is done after the deposit has been removed uct recovery (using water) on overall cleaning time of toothpaste
from a process surface, this stage is often included as part of the from a 1 m pipe. They determined that the amount of tooth-
CIP operation in industry. The purpose of this stage is to make the paste removed during product recovery was not a function of pipe
surface free of product spoilage microbes rather than to remove the Reynolds number. A similar mass fraction was removed over the
foulant. Product may be recovered before cleaning, depending on Re range 5000 to 25000. They did, however, find that prod-
the type of product, its value, and the geometry used in processing. uct recovery conditions had a profound effect on overall cleaning
This is discussed in the following section. time. The cleaning phase was conducted at the same condition:


C 2013 Institute of Food Technologists® Vol. 12, 2013 r Comprehensive Reviews in Food Science and Food Safety 129
Critical review in fouling and cleaning . . .

Table 4–Selected CIP studies of type 3 deposit.

Effect Effect of Effect of Cleaning


Soil Geometry flow temperature Effect Re chemical/pH determinant Reference
Starch (with Continuous and Local cleaning N/A, constant Re > 25000 N/A. Constant Visual, image Augustin and
phosphorescent abrupt time has a investigated. 0.5%. analysis others (2010)
tracer expansions minimum
molecules) (ID 26 mm, where the wss
expanding shows a
from 26 to 38 maximum and
mm) vice versa
Cooked SCM 316 L ss coupons Increase flow An increase in An increase in An increase from Visual, image Othman and
(sweet (square: 30 × velocity from temperature Re (6500 to 0.5 to 1.5% analysis and others (2010)
condensed milk) 30 mm long) 0.25 – 0.5 (40, 60, and 27500) NaOH did not MHFS
m/s, decrease ◦
80 C) revealed a significantly
cleaning time revealed a decrease in affect
at all linear cleaning time cleaning time
temperatures decrease in according to at higher flow
cleaning time Power law. velocities.
Egg albumin 316 L ss coupons Increase flow ◦
30 C did not Increase Re No cleaning at Visual, image Aziz (2008)
(circular: 26 less clean. (1090 to 0.1 wt% analysis, and
mm D) significant at Increase 4840) NaOH. MHFS
higher temperature, decreases Concentration
chemical con- decrease in cleaning time 0.25% to 3%
centrations. cleaning time. ◦
(at 50 C, 0.1 (at 50 ◦ C, 2.3
However, 50 to 1% NaOH). L/min)
◦ C removed At 70 ◦ C decreases
more deposit 0.1%, cleaning time.
at 1% NaOH increase Re, Most
than 70 C.◦ increase significant at
cleaning time. low flow (0.7
L/min)
WPC 316 L ss coupons Limited benefit Increase Increasing Re Limited benefit Visual, image Christian (2004)
(circular: 26 to increase temperature (1090 to to increase analysis, and
mm D) flow velocity (30 to 70 ◦ C) 4840) only concentration MHFS
horizontal ◦
at 70 C and decrease beneficial at above 0.5%.
flow cell 1% NaOH. cleaning time 0.1% NaOH.
Benefit if at all flow
increase flow rates and
at low concen- chemical con-
tration. centrations
(0.7, 1.5, 2.3
L/min, 0.1%,
0.5%, 1%
NaOH)
WPC 10 cm sections Increasing flow Wall Re 500 to 6500 N/A. Constant Thermal Gillham and
of sstubes (6 rate does not temperature investigated. 0.5%. resistance using others (1999)
mm ID 0.15 necessarily did not affect As Re MHFS and mass
mm thickness) decrease the plateau. increases
fouled in cleaning time. Increasing the cleaning time
counter It is important bulk decreases
current heat to decay temperature generally.
exchanger phase time. decreases
cleaning time.
OD = outer diameter, ID = inner diameter, ED = equivalent diameter, ss = stainless steel, MSS = mechatronic surface sensor, MHFS = Microfoil heat flux sensor, wss = wall shear stress.

0.55 m/s at 50 ◦ C. High flow velocity and low temperature in bends (see www.aeolustech.co.uk). Application of this technology
the product recovery stage revealed the fastest cleaning times. The in the food and beverage industry is also limited as the cost of
results suggested that the structure of the toothpaste film after the compressed air is considerable. However, use of air in cleaning is
product recovery stage is important in determining the overall likely to increase in the future as water becomes more precious.
cleaning time.
Product recovery can be done by pigging, in which fluid is The effect of CIP parameters on type 1 removal
expelled from a system by the “pig” which could be solid, liquid, Schlüsser (1976) compared cleaning behavior of 3 type 1 soils;
or gas. Solid pigs tend to be used in long sections of straight pipe beer, wine, and milk, illustrated in Figure 6. The products them-
work where complex geometries do not need to be navigated; for selves were not heated. The cleaning profiles of each product were
example, in crude oil pipelines to remove paraffin wax (Guo and different. Type 1 products can have a complex rheology, but are
others 2005). The use of crushed ice (with a freezing point de- often shear thinning, that is, they have an effective viscosity that
pressant) in pigging systems has been developed and researched at is a function of shear rate. The shear-thinning rheology of yogurt
the Univ. of Bristol to remove starch–water mixes (Quarini 2002). was determined by Henningsson and others (2007) who also stud-
The void fraction of the ice is controlled so that the pig can navi- ied the use of water to displace the yoghurt. For flow velocities
gate bends and T-pieces as well as straight pipe work. Application of 0.05 to 0.25 m/s, yogurt was observed and predicted to flow
of this technology in the food and beverage industries is currently as a plug. If the process was set up so that yogurt flows as a plug,
limited. The ice is expensive to make and store. A company called at changeover, the mixing zone between the 2 yogurts would be
Aeolus promotes a “Whirlwind” technology that uses compressed smaller and yield reduced losses. Prediction of the mixing zone of
air to remove soft deposits like fruit juice from pipe work with a Hershel–Bulkley material with and without wall slip at 0.19 m/s

130 Comprehensive Reviews in Food Science and Food Safety r Vol. 12, 2013 
C 2013 Institute of Food Technologists®
Critical review in fouling and cleaning . . .

surface. These are gradually eroded away according to zero-


order kinetics.

It was found in both cases that the time to remove the remaining
patches of toothpaste was the rate-limiting step in overall cleaning
time. For shampoo, Pereira and others (2009) found that flow
velocity had the biggest impact on shampoo removal from the
flow cell at the start of cleaning, less so as cleaning progressed.
Palabiyik and others (2012) found that the shear stresses induced
in the deposit during the core removal stage can affect the final
cleaning time—creation of a wavy film in the product removal
stage leads to much more rapid removal than if a smooth film is
created. The authors also found that the remaining film thickness
was independent of pipe length, suggesting that removal is uniform
throughout the pipe, as also found by Cole and others (2010).
Figure 6–Cleaning characteristics of 3 type 1 products, beer, red wine, Temperature. For cleaning of tomato paste in a flow cell, Chris-
and milk with water (Schlüsser 1976). tian (2004) found that an increase in temperature decreased the
cleaning time by a linear relationship. Both an increase in temper-
ature from 30 to 70 ◦ C and in flow velocity from 0.7 to 2.3 L/min
was also done by Henningsson and others (2007). With wall slip, decreased cleaning time. Cleaning time decreased by a factor of 6
it was predicted that the material would have a larger plug flow from the lowest flow rate and temperature to the highest flow rate
region. However, predicting the flow of a high-viscosity plug or and temperature.
wall layer is very difficult in practice. For tomato paste cleaning, it was found that cleaning time was
Flow and wall shear stress. Flow rate has an effect on the re- also correlated with Reynolds number (Christian 2004). As the Re
moval rate of type 1 materials. The rheology of tomato paste has was increased from 800 to 4800, the cleaning time (tc ) decreased
been represented by the Carreau model (Bayod and others 2008), according to a power law: tc = 2 × 106 (Re)−0.97 . R2 = 0.81.
and the cleaning behavior of tomato paste in a flow cell has been Jackson and Low (1982) found a critical Re of 6300 for cleaning
investigated by Christian (2004). At 30 ◦ C, it was found that by in- of dried tomato juice from a PHE, below which little deposit was
creasing the cleaning water flow rate from 0.7 to 1.5 to 2.3 L/min removed.
(Re 750 and 4840), the cleaning time decreased. The relationship Shampoo was rinsed at 0.14 m/s, at 31 and 51 ◦ C, by Pereira
appears linear. This was also true at 50 and 70 ◦ C. and others (2009). After the initial bulk of shampoo was removed
Shampoo (SUNSILK R color radiant, viscosity quoted as 7000
from the flow cell, it was found that the removal of shampoo
cP at 24 ◦ C) was rinsed by water from a stainless steel plate in a layers occurred faster at higher temperatures. For toothpaste, Cole
vertical flow cell by Pereira and others (2009), and they found that and others (2010) found that an increase in the water temperature
the faster the initial flow rate (in the range of 0.14 to 0.47 m/s), from 20 to 40 ◦ C decreased the cleaning time; however, increasing
the more shampoo was removed from a duct. The same effect was the temperature above 40 ◦ C did not decrease cleaning time any
found for removing toothpaste (a Hershel–Bulkley fluid with a further. The same effect may occur when rinsing shampoo; the
yield stress) from a pipe (Cole and others 2010). The effect of wall investigators did not exceed a water temperature of 51 ◦ C in their
shear stress (τ w ) in the range of 0.5 to 10 Pa on toothpaste removal experiments.
was studied. Shear stress is affected by both fluid density and Re Cole and others (2010) found that for toothpaste cleaning (from
that are both affected by temperature. Toothpaste cleaning time is various length scales and diameters), a dimensionless cleaning time,
governed by 2 removal phases by Cole and others (2010): θ c = tc u/d (where tc is the cleaning time and d is the pipe diam-
eter), could be plotted as a function of Re, as a power law model:
(i) Core removal—where most of the product is removed as a θ c = 9 × 107 (Re)−0.78 ; with a similar fit, R2 = 0.84. Palabiyik
“slug’’ of product that can be recovered. and others (2012) found that temperature had a greater effect on
(ii) Thin-film removal—where the remaining annular wall film of toothpaste film removal than flow velocity, and fitted the data.
toothpaste is removed. Design. The velocity of 1.5 m/s is the flow velocity most often
reported to clean pipe lines effectively in industry CIP (EHEDG
More recently, 3 phases were defined by a further investigation
1992). This is, however, anecdotal with no theoretical justifica-
of the effect of product recovery on cleaning of toothpaste using
tion (Changani and others 1997; Tamine 2008). In industrial pipe
water (Palabiyik and others 2012):
systems, there are, however, more complicated geometries such as
(i) Core removal—the first few seconds (a time comparable to bends, valves, and T-pieces. It raises the question: does increasing
the residence time of the fluid in the system) where approx- the flow velocity decrease the cleaning time of other geometries?
imately half the product mass is removed and the remaining This gives a better indication of the effect of flow on the cleaning
toothpaste coats the pipe wall. Also called the product re- time of a whole system.
covery stage. Jensen and others (2007) filled a variable depth “upstand” or
(ii) Film removal—further product is removed up to about “downstand” (also called a T-piece) made from glass with com-
1000 s according to a process that is 1st order in deposit mercially available mustard and rinsed with ambient water. The
weight/thickness, leaving a thinner but continuous film of geometry used in the study is shown in Figure 7 and is in the
toothpaste remaining on the pipe wall. downstand position. The downstand depth was tested at 4 and 6
(iii) Patch removal—greater than 1000 s, the continuous film is cm. The flow velocity was increased from 1 to 1.88 m/s to define
broken up and only patches of toothpaste are left on the the effect on cleaning the T-piece. Jensen and others (2007) found


C 2013 Institute of Food Technologists® Vol. 12, 2013 r Comprehensive Reviews in Food Science and Food Safety 131
Critical review in fouling and cleaning . . .

test, the apparatus is filled with sour milk and/or spores. An area
“difficult to clean” is defined as an area that produces yellow agar
in 3 consecutive tests (EHEDG 1992). Yellow agar shows the
presence of spores. The study revealed that the valve was easier to
clean than the radial flow cell (detailed by Jensen and Friis 2004).
The study predicted that a critical wall shear stress of 3 Pa was
necessary in both systems to ensure cleaning; however, areas of
extremely low wall shear stress and some areas of wall shear stress
higher than 3 Pa had spores remaining. The authors concluded
that wall shear stress was not the only factor governing cleaning
in this case. As spores are more likely a type 2 soil, this conclusion
seems logical.
Bénézech and others (2002) rinsed spores in custard from a
progressive cavity pump (a type of positive displacement pump)
using a standard CIP operation in 2 configurations (i) with an axial
exit pipe, where custard was pumped out of the top of the pump
Figure 7–Downstand geometry used for investigating the influence of
different flow rates during CIP (flow was from left to right) (from Jensen body on the same axis as entry, and (ii) with a tangential exit pipe,
and others 2007). custard was pumped out of the body at the side off the axis of
entry. The CIP consisted of a prerinse at 0.5 m/s for 6 min; 0.2%
NaOH rinse at 1.5 m/s, 60 ◦ C, for 10 min; intermediate rinse at
that: 0.5 m/s for 6 min; 0.2% HNO3 rinse at 1.5 m/s, 60 ◦ C, for 10
min; and final rinse at 0.5 m/s for 6 min. The group found that in
(i) Increasing the flow velocity increased removal rate. How-
the tangential setup, all parts of the pump were cleaned to the same
ever, the authors suggested that this was more likely due
number of CFU/cm, approximately 10 CFU/cm2 . The authors
to greater acceleration of the water at 1.88 m/s into the
defined a high level of hygiene as counts less than 18 CFU/cm2 .
T-piece. At the lower flow velocities, flow had not fully
In the axial set up not all components were cleaned to the same
developed before entering the T-piece.
level. There was an increased number of CFU/cm2 in the pump
(ii) Some areas of the T-piece were harder to clean than others.
body and gaskets (>18 CFU/cm2 ).
The position in the downstand most difficult to clean was
To clean tanks, spray devices typically called cleaning heads are
always located in the same position (see Figure 8, shown as
used. The design of a cleaning head is of paramount importance
a downstand).
to be effective in cleaning. There are 2 main choices:
As expected, the top of the downstand was difficult to clean.
(i) Static cleaning heads—These devices spray cleaning fluid onto
However, an additional area located on the downstand pipe was
the tank surface from a fixed position. The effectiveness of
always the last part to be cleaned in all the experiments, regardless
the cleaning head depends on cleaning fluid flow rate and
of velocity. Jensen and others (2007) used computational fluid
the size and pattern of the holes.
dynamics (CFD) simulations to predict the wall shear stress in the
(ii) Dynamic cleaning heads—These devices spray cleaning fluid
4-cm downstand. Their CFD findings are illustrated in Figure
onto the tank surface using larger pressures, around 5 Bar
8(A) to 8(C) where blue is low wall shear stress (0 Pa) and red is
(resulting in large wall shear stresses and direct impact force),
high wall shear stress (5 Pa). As the flow velocity was increased,
and rotation to ensure full vessel coverage. The effectiveness
the blue area decreased in size. Within these simulations, the area
of the cleaning head depends on the cleaning fluid pres-
most difficult to clean, the center of the downstand, is identified.
sure/flow rate to ensure that the preprogrammed pattern is
Increasing the flow rate does not improve cleaning of this area.
achieved.
The wall shear stress achieved at this position is low at all 3 flow
velocities. The other areas hardest to clean are circled. Examples of commercially available cleaning heads of both types
Jensen and others (2007) examined the effect of pulsed flow in are shown in Figure 9. Increasing the impact force of a jet stream
the downstand. They found that pulsing flow only affected the of fluid onto a surface can overcome large deposit hydration times
cleaning time of the 4-cm-depth T piece, not the 6-cm-depth and reduce cleaning times. The fraction effect of time, physical
T piece. They compared cleaning at 1 m/s (v1 ) and 2 m/s (v2 ) action, temperature, and chemical action delivered to the tank by
and pulsing at 15 s (p1 ) and 30 s (p2 ). The cleaning time of the a static cleaning head (spray ball) and a dynamic cleaning head
4-cm downstand was longer when rinsed at 1 m/s than when (high-pressure cleaning head) are given in Figure 10 (Tamine
the flow was pulsed. However, rinsing the downstand at 2 m/s 2008). For spray ball cleaning, time is required to achieve deposit
gave the quickest cleaning time. The authors concluded that at removal. Cleaning time is required to achieve product removal
turbulent Re, the area of the recirculation zone in the T-piece did using a static head, and mechanical action is required to achieve
not change. A recirculation zone is typically located after a pipe product removal using a dynamic cleaning head. Dynamic heads
expansion and depends on Reynolds number and the expansion enable cleaning behaviors that are less reliant on contact time
ratio. At lower Re (less than 10000 in this case), the length of the with the chemical at high temperature. A type 3 soil could be
recirculation zone may change; hence, cleaning times are shorter cleaned in a similar time as a type 1 soil. It should, however, be
for pulsed flow at 1 m/s than using constant flow at 1 m/s. noted that impingement jets from a rotary device or from many
Jensen and Friis (2005) used CFD simulations to predict the small jets in a static device may cause corrosion problems due to
cleanability of a mix proof valve fouled with B. stearothermophilus “rouging,” from small iron particles worn from the orifices of
spores in accordance with the EHEDG standard cleanability test the thin walled static spray devices that then deposit on the tank
(EHEDG 1992; Timperley and others 2000). In the EHEDG wall.

132 Comprehensive Reviews in Food Science and Food Safety r Vol. 12, 2013 
C 2013 Institute of Food Technologists®
Critical review in fouling and cleaning . . .

A B C

Figure 8–CFD simulations of the flow field in 4 cm downstand T piece at (A) 0.5 , (B) 1, and (C) 2 m/s. Blue is low wall shear stress (0 Pa) and red is
high wall shear stress (5 Pa). White represents wall shear stress in excess of 5 Pa. Water enters the section from the right and exits the T section on
the left represented by the arrow in (a) (adapted from Jensen and others 2007).

A B C

Figure 9–Commercially available (A) spray ball static device (GEA, Warrington, U.K.), (B) rotary spray head dynamic device (Alfa Laval, Minworth,
U.K.), and (C) rotary jet head dynamic device (Alfa Laval, U.K.).

Morrison and Thorpe (2002) defined the wetting rate at the als, and lipids. The deposits formed during wort boiling are solid
mass flow rate (kg/s) required to completely wet a surface of and dissimilar to the process stream (Tse and others 2003).
width W (in meter). Wetting rates achieved by single jets from Membrane cleaning. There are many types of filtration pro-
a spray ball were 0.1 to 0.3 kg/ms. The act of removing deposit cesses in food and beverage manufacturing operations. The fouling
from a vessel involves initial wetting and subsequent softening of membranes alters permeability and selectivity and can be char-
(or dissolution) of the deposit, followed by complete removal by acterized by increased pressure differential and decreased mem-
further impingement. Morrison and Thorpe (2002) measured the brane flux. Membranes used in the food and bioprocess industries
dimensions of the wet area by the impaction of single water jets include reverse osmosis (RO), nanofiltration (NF), ultrafiltration
onto a sheet of painted acrylic for a range of pressures and distances (UF), and microfiltration (MF) (Cui and Muralidhara 2010). In
from spray balls of different sized holes. They found that if the the brewing industry, beer is clarified using MF in which yeast
jet directly impacted the area to be cleaned, then this area was readily fouls the membranes. Membrane cleaning is complex as it
cleaned within 60 s. The point of impact was smaller than the is necessary to both remove the surface layers and open the pores
total area being wetted; however, certain areas were not cleaned in the structure—this must be done without the cleaning agent
by the spray ball. The width of the falling film from the point damaging the material.
of impact remained the same size throughout rinsing. Jet breakup Güell and others (1999) found that when yeast cells were present
was observed at 45 ◦ C, which increased the distribution of the on cellulose acetate membrane (CAM) as a layer (yeast cake), the
jet and cleaned a larger area. An interesting model for the flow yeast was believed to have formed a secondary membrane. In-
behavior of jets is given by Wilson and others (2012). creasing the thickness of the yeast cake reduced permeate flux and
protein transmission through the membrane. Increasing the yeast
concentration in the feed solution resulted in lower fluxes and
The effect of CIP parameters on type 2 and type 3 deposit protein transmission through the CAM. Hughes and Field (2006)
removal discussed the fouling of MF and UF membranes with yeast at sub-
Type 2 deposits can be viscoelastic, temperature-dependent, critical fluxes where fouling is negligible. For the MF membrane,
and/or time-dependent (Rao 1999). Type 3 deposits tend to be the rate of fouling increased with increasing feed concentration,
thermally induced and precipitate from the process stream onto the increasing membrane pore size, and decreasing shear stress. The
heat exchanger surface over time. For example, wort is a complex UF membrane could not be cleaned effectively.
fluid with several components that change structure and solubility Mores and Davis (2002) examined the effect of pulsing flow
upon heating, including carbohydrates, proteins, vitamins, miner- through a CAM to clean it. They found that the flux increased


C 2013 Institute of Food Technologists® Vol. 12, 2013 r Comprehensive Reviews in Food Science and Food Safety 133
Critical review in fouling and cleaning . . .

less steel membrane at 50 ◦ C, 1.67 m/s. Increasing the CFV from


1 to 6 m/s decreased fouling resistance of the ceramic membrane
and gave the least amount of fouling present on the membrane
after 20 min.
Water rinsing of hard surfaces. For the type 3 deposits WPC
and egg albumin, Christian (2004) and Aziz (2008) found that
neither deposit was removed with water rinsing at the temperatures
and flow velocities investigated; 30 to 70 ◦ C and 0.7 to 2.3 L/min.
The authors determined that chemical action was required for their
removal.
Guillemot and others (2006) rinsed rehydrated Saccharomyces
cerevisiae cells from stainless steel in a flow cell over a wall shear
stress range of 0 to 80 Pa. They found that as the wall shear
stress was increased, the number of cells remaining on the steel
decreased. However, only a 10% reduction in the number of yeast
cells was achieved in this range of wall stress. Goode and others
(2010) rinsed aged yeast slurry from stainless steel coupons using
water in a flow cell, and they found that increasing the flow velocity
did not significantly affect the amount of deposit removed from
the surface at ambient temperature; this was over a wall shear
stress range of 0 to 1.24 Pa. Rinsing removed around 50% of the
deposit area. Goode and others (2010) also found that increasing
the temperature of the water rinse removed more deposit up to
50 ◦ C with flow rate having a negligible effect. However, at 70

C, decreased removal efficiency was observed, particularly at the
highest flow velocity, 0.5 m/s.
The yeast was aged at different temperatures and for different
times in the work by Guillemot and others (2006) and Goode and
others (2010); 20 and 30 ◦ C and 1 h and 5 d, respectively. The cell
concentration was also different at 0.0065 g/mL for Guillemot and
others (2006) and 1 g/mL for Goode and others (2010). These
findings suggest that fouling conditions dictate cleaning behavior,
as already found from milk fouling (Changani and others 1997).
Chemical effects on the cleaning of type 2 deposits. Various
authors have investigated the removal behavior of bacterial spores
from stainless steel. The effect of shear on adhesion has been
studied using devices such as the radial flow cell (Detry and others
Figure 10–The fractional importance of different factors: time, coverage, 2007, 2009) that can, when used correctly, allow ranges of shears
physical action (impact), temperature, and chemical action (chemistry) to be studied. Le Gentil and others (2010) cleaned Bacillus cereus
required for effective tank cleaning by (A) spray ball and (B) rotary jet spores from 316 L stainless steel pipes using 0.5% (w/w) NaOH
head (adapted from Tamine 2008). at 60 ◦ C at 2.2 L/min. The test was carried out over 30 min.
As the cleaning time increased, the number of spores decreased
as expected. In the first 10 min, up to 70% of the spores were
with increasing shear rate, back pulse pressure, and back pulse removed, less so in the remaining 20 min. Lelièvre and others
duration. At higher shear rate and back pulse pressure, multiple (2002) investigated the removal of B. cereus spores from 304 L
short back pulses were more effective in cleaning the membrane. stainless steel pipes, similar in length and diameter to the pipes
At low shear rate and back pulse pressure, fewer longer back pulses used in the study by Le Gentil and others (2010) and 0.5% (w/w)
were more effective. Longer, weaker back pulses led to the highest of NaOH was used at 60 ◦ C to rinse the pipe. In this study, the
recovered fluxes. effect of flow velocity and temperature was investigated over a 30
Shorrock and Bird (1998) fouled a MF membrane (hydrophilic min clean. The researchers found that cleaning at 60 ◦ C removed
polyethersulfone, 0.1 μm pore diameter), with yeast cake. Water more spores than rinsing at 20 ◦ C at each 5-min time interval, at
rinsing was found to remove most of the deposit and an increase the same flow velocity of 1.97 m/s. They found that increasing the
in temperature from 30 to 60 ◦ C was found to decrease fouling flow velocity from 1.61 to 3.29 m/s (τ w = 17.45 to 68.95 Pa) at
resistance (at 0.74 m/s cross-flow velocity (CFV)). At 40 ◦ C, us- 60 ◦ C decreased the number of attached spores in the first 5 min.
ing NaOH as optimum concentration, there was optimum flux However, after this time, the contact time was more important
through the membrane, 0.01% to 0.025%. Formulated sodium hy- in removing the spores. The increased acceleration at higher flow
droxide solution was found to restore membrane flux completely. rates may be controlling the number of spores removed in the first
Cleaning of MF membranes with WPC was considered by Bird 5 min of cleaning, as found by Jensen and Friis (2004).
and Bartlett (2002) using a flat plate stainless steel membrane and Bremer and others (2006) investigated the effect of alkali rinses
by Blanpain-Avet and others (2009) using a tubular ceramic mem- and acid rinses (formulated and nonformulated) on removing a
brane. An optimum alkaline detergent concentration of 0.02% biofilm generated by recirculating skimmed milk powder in a CIP
NaOH was found to give maximum flux after cleaning of the stain- skid for 18 h in 15 mm stainless steel tubes. There were a number

134 Comprehensive Reviews in Food Science and Food Safety r Vol. 12, 2013 
C 2013 Institute of Food Technologists®
Critical review in fouling and cleaning . . .

of conclusions: when moving from laminar to turbulent flow. Disruption of the



boundary layer is further discussed in Section 4.1. Generally in-
(i) Rinsing with 1% NaOH (for 10 min, 65 C, 1.5 m/s) creasing the temperature decreases the cleaning time. Gillham and
followed by 1% nitric acid (for 10 min, 65 ◦ C, 1.5 m/s) others (1999) found that removal of whey protein deposits from
reduced the number of cells to a similar level than that found stainless steel pipes was strongly dependent on temperature (less so
after rinsing with only NaOH (at the same conditions). the swelling phase).
(ii) Formulated detergents (with surfactants, chelating agents, SCM is an intermediate in the manufacture of some confec-
and sequestrants) decreased cell numbers to the same level tionary products, made by evaporating water from milk and adding
as rinsing with NaOH (at the same conditions). sugar to lower the water activity of the product. SCM has 70%
(iii) Addition of a surface-active agent to the caustic solution sig- to 74% total solids of which 40% to 45% is sucrose (Fisher and
nificantly reduced the number of cells compared to standard Rice 1924) leaving 29% to 30% milk solids. In the study of Oth-
CIP (NaOH and nitric acid in (i)). man and others (2010), SCM was cooked for 4 h at 85 to 90 ◦ C
(iv) Nitric acid with surfactants removed significantly more cells on stainless steel coupons and washed by chemical cleaning in a
than just nitric acid. flow cell. It was found that increasing the flow velocity from 0.25
(v) Addition of a sanitizer step after CIP did not significantly to 0.5 m/s decreased the cleaning time at all temperatures. An
reduce viable bacteria numbers. increase in temperature from 40 to 80 ◦ C decreased the clean-
This suggests that the concentration of the alkali, the flow velocity, ing time linearly. Interestingly, the authors found that increasing
and the temperature could be optimized to give the most efficient the NaOH concentration from 0.5% to 1.5% did not significantly
cleaning regime where all cells can be removed. affect the cleaning time at each temperature. This agrees with
Goode and others (2010) investigated the effect of chemical on findings for WPC cleaning that quote 0.5% NaOH as the opti-
yeast removal from stainless steel coupons in a flow cell using 2% mum concentration. It was the increase in temperature rather than
Advantis 210 (1% NaOH equivalent). They found that increasing the increase in chemical concentration that decreased cleaning
the temperature from 20 to 70 ◦ C decreased the cleaning time. time.
An increase in flow velocity at 50 and 70 ◦ C from 0.26 to 0.5 m/s Cleaning time was plotted compared with Re for SCM at 1%
also decreased the cleaning time; however, at 20 and 30 ◦ C, an NaOH (40, 60, and 80 ◦ C) by Othman and others (2010) and at
increase in flow velocity from 0.4 to 0.5 m/s did not significantly 0.1%, 0.5%, and 1% NaOH (at 30, 50, and 70 ◦ C) for WPC by
decrease cleaning time. Christian (2004) as shown in Figure 11(A) and 11(B). For WPC,
Chemical effects on the cleaning of type 3 deposits. The effect the range of investigated Re was around 800 to 4840. There were
of chemical cleaning of WPC from milk has largely been char- separate groups of data at each temperature that could not be
acterized in the literature as uneven. The cleaning process has 3 plotted on one master curve. This suggests that temperature was
distinct phases seen by many independent researchers, for exam- the dominant parameter in controlling cleaning time. Christian
ple, by Bird (1992), Gillham (1997), Grasshoff (1997), Tuladhar (2004) concluded that an increase in Re was only beneficial to
(2001), and Christian (2004): cleaning time at low concentration. Jennings and others (1957)
suggested the existence of a threshold Re of 25000 for cleaning a
(i) Swelling—alkali solution contacts the deposit and causes pipe surface of dry milk deposit before an increase in Re resulted
swelling, forming a protein matrix of high void fraction. in increased cleaning rate. For SCM, the Re range investigated was
(ii) Erosion—uniform removal of deposit by shear stress forces much higher, from 6500 to 27000. All the data collapsed onto one
and diffusion. There may be a plateau region of constant clean- curve. As the Re increased, the cleaning time decreased, suggesting
ing rate, but this depends on the balance between swelling that Re was the dominant parameter controlling cleaning time.
and removal. Othman and others (2010) did find, however, that the effect of
(iii) Decay—the swollen deposit is thin and no longer uniform Re on cleaning time became less significant as the temperature was
so that removal of isolated islands occurs by shear stress and increased. Gillham and others (1999) found that tc was proportional
mass transport. to Re−n , where n was in the range of 0.2 to 0.35 for 0.5% NaOH.
Many authors quote 0.5% NaOH to be optimal for WPC removal For SCM at 1% NaOH, tc was again proportional to Re−n where
from stainless steel, although the existence of cleaning optima has n = −1.28 (R2 = 0.92).
not been categorically proved in all cases. Bird and Fryer (1991) The cleaning of egg albumin was characterized by Aziz (2008).
found that increasing the NaOH concentration to 2% can pro- Generally, an increase in temperature decreased the cleaning time.
duce a deposit with a less open (dissolved) structure than at 0.5%, However, at 1% NaOH, cleaning time was faster at 50 ◦ C than
thus lengthening the swelling phase—Yoo and others (2007) and at 70 ◦ C. Deposit was not removed at 30 ◦ C at any flow velocity
Saikhwan and others (2010) explained the processes that underpin or NaOH concentration investigated. An increase in NaOH con-
this observation. Plett (1985) reported that a maximum cleaning centration from 0.25% to 3% NaOH decreased the cleaning time
rate occurs when cleaning with detergent. The contribution of (at 50 ◦ C, 2.3 L/min); however, increasing the flow rate had a less
flow rate is hard to determine in chemical cleaning because both significant impact on cleaning time at higher chemical concentra-
shear stress imposed on the deposit and mass flow to the deposit are tion. The author concluded a high temperature, mid to high flow
dependent on the flow rate. In general, the higher the flow rate, velocity, and mid-range chemical concentration appeared to be the
the shorter the cleaning time. Timperley and Smeulders (1988) optimum, similarly to WPC cleaning optima. For egg albumin,
found that the cleaning time of a PHE decreased with increasing the range of Re investigated was 1090 to 4840. There were sepa-
flow velocity from 0.2 to 0.5 m/s. There are arguments supporting rate groups of data at 50 and 70 ◦ C that could not be plotted on
higher flow rates that create turbulent conditions. This is because one master curve. This suggests that temperature was the dominant
turbulent conditions are known to make the flow patterns of the parameter in controlling cleaning time similarly to WPC.
microscopic boundary layer unstable. However, Bird and Fryer In Christian (2004), WPC cleaning experiments were con-
(1991) found that there was no significant change in cleaning rate ducted using 0.5% NaOH at 30, 50, and 70 ◦ C and 0.7, 1.5, and


C 2013 Institute of Food Technologists® Vol. 12, 2013 r Comprehensive Reviews in Food Science and Food Safety 135
Critical review in fouling and cleaning . . .

Figure 12–Stationary flow and oscillating components of flow, wos


symbolizes an oscillating fluid movement, wos,max is maximum oscillating
fluid velocity, wstat is mean fluid velocity, and ωt is frequency of oscillation
(Augustin and others 2010).

an oscillating fluid movement, wos , is superimposed, as illustrated


in Figure 12.
The intensity of a superimposed pulsation (wos,max + wstat ) can be
quantified using waviness, W, the ratio of the maximum oscillating
(wos,max ) and the stationary or mean flow velocities, w:
wos,max
W= (4)
w
w for an oscillation interval is defined by Augustin and others
Figure 11–Re compared with visual cleaning time of (A) SCM at 40, 60, (2010) as:
and 80 ◦ C using 1% NaOH (from Othman and others 2010) and (B) WPC
at 30, 50, and 70 ◦ C, using 0.1%, 0.5%, and 1% NaOH (from Christian

1 tos
2004). Shading denotes the flow temperature: Black: 30 ◦ C, Gray: 50 ◦ C, w= w (t ) d t with w(t ) = wstat + wos
and Open: 70 ◦ C. tos 0 (5)
= wstat + wos,max . sin(wt )
2.3 L/min. Rd (the fouling resistance) is a measure of resistance A higher value of waviness is believed to result in a separation
to the flow of heat to the sensor. Rd measured at the same flow of the viscous sublayer and the formation of eddy currents; flow
rate reduced Rd more rapidly as the temperature was increased reversal is critical. This effect can decrease the thickness of the
from 30 to 70 ◦ C. An increase in flow rate from 1.5 to 2.3 L/min laminar sublayer at the pipe surface when applying a turbulent
revealed similar Rd profiles, suggesting that temperature dictated flow. The temporary maximum velocity, wos,max , can occur near
the cleaning time in this case. the pipe wall resulting in large shear rates and high wall shear
For all type 3 deposits detailed here, temperature seems to be the
stresses. Bode and others (2007) characterized a linear decrease
dominant contributor to cleaning time at both low and high flow of cleaning time with increasing waviness from 1 to 5. Below a
velocity and low and high concentration. Reaction rate, solubility, waviness of 1, there is limited flow reversal in the pipe and the
and possible phase transitions (such as in fats) will all be affected
cleaning time increases. Augustin and others (2010) compared the
by temperature. cleaning rate of deposit at a flow velocity of 1 m/s (Re 25000),
where waviness was 0, and pulsing the flow, where waviness was 1.
Novel Cleaning Approaches They found that the amount of deposit in the pipe became asymp-
There have been various methods reported in the literature to totic at 4 min using the pulsed-flow regime and 6 min using the
improve cleaning. stationary flow regime. Augustin and others (2010) have validated
CFD models accompanying their cleaning data. Phosphorescent
Increasing boundary layer disruption zinc sulfide crystals were used as an optical tracer enabling accurate
Various authors have considered pulsing flow in pipes to en- characterization of flow patterns.
hance wall shear stress at lower average flow velocities to enhance
cleaning. Gillham and others (2000) showed that pulsing flow at Alternative cleaners to reduce environmental impact
a relatively low frequency (2 Hz) enhanced cleaning of a tubu- The efficacy of enzymes in cleaning has been investigated over
lar heat exchanger compared to the same steady-flow velocity. A the last decade. Grasshoff (2002) investigated the efficacy of Sav-
pulsed flow creates high periodic accelerations of the liquid flow. inase, a protease (optimum temperature 50 ◦ C, pH 9.5), in cleaning
The directional change in the flow can increase mass transfer of a milk pasteurizer following a 15-min acid wash. Increasing the
the cleaning fluid to the surface, thus decreasing cleaning times. concentration of the enzyme from 0.0025% to 0.05% at 60 ◦ C
A pulsed flow is characterized by a stationary base flow on which showed that residual soil was removed faster. In the plant, the

136 Comprehensive Reviews in Food Science and Food Safety r Vol. 12, 2013 
C 2013 Institute of Food Technologists®
Critical review in fouling and cleaning . . .

heat exchanger was opened and inspected after 45 min of enzyme spectively. The micromanipulation technique employs controlled
cleaning and 6 min of water rinsing (30 L/min). The interior strain parallel to the deposit, where a T-shaped probe is pulled
surfaces were clean. Microbiological product samples collected across a horizontal circular plate at a constant height, removing
showed no indication of microbial or enzyme contamination after the fouling deposit by a shoveling action. The FDG was devel-
CIP. oped to measure the thickness of soft-solid meso- and macrolayers
The use of commercial enzymes to clean UF membranes has on surfaces. The surface is immersed in a Newtonian liquid and
been discussed by Petrus and others (2008) and Allie and others a nozzle is brought close to the surface. A suction pressure dif-
(2003). Petrus and others (2008) used proteases to clean proteins ferential is applied so that liquid is drawn into the nozzle. The
(BSA and β-Lg) and defined an optimum concentration of 0.1% gauging suction is increased until deformation is observed. Defor-
for 60 min. The enzyme deposited on the membrane when rinsed mation of the deposit results in changes in h, the distance between
for longer than 60 min. Allie and others (2003) used proteases and the deposit and the gauge tip, which can be readily detected.
lipases to clean abattoir effluent. Up to 55% of fouling was removed The 2 techniques showed similar trends and complementary phe-
with lipases, and up to 70% by using lipases and proteases and a flux nomenological detail when removing tomato paste in parallel stud-
recovery of up to 100%. However, to apply enzyme cleaning in ies (Hooper and others 2006). As the baking time increased, the
industry enzyme dosage, process control, and the overall economic adhesive strength increased, and as the hydration time increased,
burden need to be considered. the adhesive strength decreased.
Orgaz and others (2006) tested the efficacy of fungal enzymes at The FDG technique has been used to study the effect of alka-
breaking Pseudomonas fluorescens biofilm bonds. Out of the tested line cleaning chemical concentration (0.3% to 2%), temperature
enzymes, a T. viride enzyme (cultured on pectin) was most ef- (20 to 50 ◦ C), and velocity (0.03 to 0.3 m/s, Re: 500 to 10000)
fective at removing the biofilm by 84% (± 2%). The enzyme on whey protein (Tuladhar and others 2002), the effect of drying
was cellulose-pectinesterase-based. The least effective enzyme re- time on tomato paste (Chew and others 2004), and the effect of
moved 19% (± 6%) of the biofilm and was mostly cellulose. temperature and pH on gelatin film swelling (Gorden and oth-
ers 2010). The micromanipulation technique has been used to
Other studies related to cleaning behavior study the strength of P. fluorescens biofilm with growth medium
Various authors have quantified parameters related to cleaning velocity (Chen and others 1998), the effect of baking, hydration
behavior by other methods rather than in a flow cell or pilot plant. time, and temperature on tomato paste (Liu and others 2002), and
Some techniques used to infer cleaning data are discussed here. surface energy and cut height on tomato paste (Liu and others
Deposit shear. Simões and others (2005) fouled stainless steel 2006). The same study considered the effect of drying/aging and
cylinders with P. fluorescens biofilm in a bioreactor. Three cylinders cut height on bread dough, egg albumin, and whey protein. The
were rotated at 500, 1000, 1500, and 2000 min−1 sequentially effect of ovalbumin concentration, temperature, and NaOH con-
(ReA 2400 to 16100) for 30 s each in phosphate buffer to assess centration on egg albumin adhesion is presented by Liu and others
the effect of rotation speed on biofilm removal. There was an (2007). Tomato paste, bread dough, and egg albumin deposits have
optimum ReA of 8100 where 45% of biofilm was removed. Either been found to have a lower adhesive than cohesive strength, while
side of this ReA removal was less and similar at around 15%. whey protein was found to have lower cohesive than adhesive
Cylinders were also submerged in different chemical solutions and strength.
rotated at 300 min−1 for 30 min and then were rotated at 500, Atomic force microscopy (AFM) has also been used to study
1000, 1500, and 2000 min−1 sequentially in phosphate buffer deposit–surface interactions. In this technique, a cantilever con-
to test cleaning effectiveness. NaOH and sodium hypochlorite tacts the deposit and the force of detaching the cantilever is de-
(NaClO) representing a CIP detergent and CIP sterilant were termined. Beech and others (2002) demonstrated the effective use
investigated. Irrespective of ReA , NaOH and NaClO removed a of AFM for studying the adhesion of bacteria to a wide range
similar percentage of biofilm at the same concentrations (50, 200, of surfaces. Bowen and others (2001) used AFM to study yeast
and 300 mg/L); however, at the highest concentration, 500 mg/L, cell detachment from hydrophobic and hydrophilic silicate sur-
NaOH removed approximately 5% more biofilm. faces with and without protein. The authors found that as yeast
Demilly and others (2006) characterized the removal of yeast cells aged, adhesion force changed. Also, cells in the stationary
cells as a function of wall shear stress from different stainless steel phase adhered most strongly to the surface. Surface contact was
surfaces using a radial flow chamber, in which flow trajectory demonstrated to be important where after 5 min, adhesion force
was from the center of the plate outwards toward the edges of increased. Elofsson and others (1997) used AFM to study the cov-
the plate. This is similar to the impingement of a water jet on erage of β-Lg and WPC to mica sheets at different surface loadings
a surface. Radial flow investigations may relate to the action of in the range of 0.03 to 3 kg/cm2 . The use of AFM enabled the au-
a cleaning head in a tank, to provide a microscale method of thors to distinguish between different states of protein aggregation
investigating water jets, and deposit removal. The steel surfaces at the submicrometer level.
were micropolished and electrochemically etched with different
sizes of grain and depths. The number of cells remaining from Measuring Cleaning
the original number of cells was defined. Interestingly, the authors Ensuring good performance of the cleaning process is vital in
found a threshold shear stress above which cells were removed maintaining reliable product shelf life and quality. Figure 1 demon-
regardless of topography; that detachment of yeast cells was faster strates implementation of a CIP standard in industry, revealing
from etched steel than mirror-polished steel. that CIP measurement and control are fundamental to ensuring
Deposit deformation and strength. Two methods have been in- hygiene on a daily basis. Processes used to establish and run CIP
dependently developed at Birmingham and Cambridge to deter- operations include:
mine the strength and deformation behavior of soft-solid fouling
layers on hard surfaces immersed in liquid: the micromanipulation (i) Validation—determining the right method of cleaning and
technique and the fluid dynamic gauging (FDG) technique, re- setting it as a standard; it is done before implementation of


C 2013 Institute of Food Technologists® Vol. 12, 2013 r Comprehensive Reviews in Food Science and Food Safety 137
Critical review in fouling and cleaning . . .

drop initially increased across a PHE during cleaning. When the


cleaning chemical came into contact with the deposit, the deposit
swelled and further increased the pressure drop through the pro-
cessing plant before cleaning commenced. This effect has been
well characterized in pulsed-flow cleaning by Christian and Fryer
(2006). Robbins and others (1999) also used pressure drop across a
PHE to characterize the fouling nature of milk and whey protein
fouling.
Van Asselt and others (2002) demonstrated online monitoring
of a dairy evaporator CIP by conductivity and pH. The chemical
and water interfaces were clear. Conductivity can also be used to
indicate product–water interfaces in nonchemical cleaning. Fickak
and others (2011) measured conductivity during water rinsing (at
0.01 m/s) of a heated rod fouling with heat-induced whey pro-
tein gel (HIWPG). The conductivity was seen to increase to 1200
μS/cm in the first 100 to 200 s and decrease thereafter to 200
μS/cm in around 1000 s. The decrease in conductivity was a
smooth slope. Fickak and others (2011) also used turbidity at the
outlet of the fouled rod to indicate removal behavior during the
prerinse. The fouled rod was rinsed until the turbidity measure-
Figure 13–Potential locations of online CIP measurement techniques that
could be used in a process line. ment was at 0.5 to 1 Nephelometric Turbidity Units (NTU),
indicating drinking water quality. The turbidity increased at the
onset of the prerinse to approximately 28 NTU in the first 200 s
a new method and after alterations in an existing operation. and decreased thereafter. At 1200 s, however, the turbidity value
It should always be up-to-date. increased from 3 to 7 NTU before decreasing further. The tur-
(ii) Verification—are the results correct and accepted? Checks bidity profile contained numerous jagged peaks that may suggest
that the system behaves in the predetermined and expected the deposit removal was nonuniform. The profile is quite dif-
way, after validation. ferent from conductivity. Van Asselt and others (2002) tested an
(iii) Monitoring—continuous monitoring of specific points of a inline calcium probe (CHEMFET) during cleaning of a tubu-
process determining that the process is under control, after lar heat exchanger against offline measurements of calcium (g/L)
verification. and pH. During the prerinse, no difference in the online mea-
When defects in the system have been identified, appropriate ac- surement was detected. Offline measurements revealed a decrease
tion should always be taken as detailed in Figure 1. Typical CIP in calcium from 0.5 to 0 g/L in approximately 3 min and an
monitoring tools include visual detection methods, microbial enu- increase in pH from approximately 6 to 8 in 6 min. Conduc-
meration of CIP rinse water, and in-line probes that measure tem- tivity that may have been measured during the prerinse was not
perature and the proportion of hydrogen ions (pH) or electrolytes presented.
(conductivity) in the cleaning fluid. The conductivity of acid and Van Asselt and others (2002) also measured turbidity and con-
alkali is different to that of water so that the chemical phases of ductivity online and offline during cleaning of a dairy evaporator.
cleaning can be clearly identified and the chemical recovered. The The evaporator was not prerinsed with water. The conductiv-
first step in optimizing CIP is ensuring that the used monitoring ity revealed, as expected, an increase during chemical phases and
techniques are giving reliable data. Yang and others (2008) de- a decrease during water phases. Turbidity determined in-line and
scribed the importance of forecasting models in the approach to offline gave different deposit removal profiles. For example, during
optimize cleaning. Data obtained at the beginning of CIP (where the second alkaline rinse online turbidity revealed 2 considerable
the confidence limit of variation is high) are used to predict the end peaks. This was also seen at the onset of acid to a lesser extent.
point of cleaning (where the confidence limit of variation is low). Offline measurements did not reveal these peaks. The offline tur-
Online application of such a tool would be valuable in optimizing bidity samples taken during CIP were analyzed at a later stage
CIP. A range of measurement devices have been considered in which the authors think altered the particle size of the samples,
the literature and have been used to assess either the cleaning fluid and thus the turbidity measurement. Particles of different sizes
(bulk measurements) or the surface. Figure 13 illustrates where in a scatter light differently, for example, as a function of smaller par-
process of line bulk and surface measurements, CIP measurements ticle size (Clauberg and Marciniak 2009). However, nonproduct
may be taken. Flow, temperature, conductivity, and pH are bulk substances (such as air bubbles and detergent) can also absorb light
measurements. giving misleading data.
CellFacts (Coventry, U.K.) has a patented offline sampling tech-
Online bulk measurements nology which they have demonstrated determines particle size
Online bulk measures monitor the fluid during cleaning. Typ- and density in rinse water samples. This was demonstrated in case
ically temperature, flow, pressure, conductivity, and pH are mon- studies on their website (http://www.cellfacts.com/Cleaning-In-
itored during CIP to ensure that the process is under control. Place.php). The company used this technology to determine par-
Pressure drop across a system (P) is defined as: P = PI − ticle size and density during CIP of a beer maturation vessel (Scot-
PO , where PI and PO are the inlet and outlet pressures. PO in- tish & Newcastle Breweries 2008). Malvern Instruments (Malvern,
creases with time during fouling and decreases during cleaning U.K.) have made commercially available technologies that could
generally. The rate of change of pressure can be used as a mea- be used to monitor particle sizes and counts online during CIP of
sure of cleaning. Fryer and others (2006) illustrated that pressure a process line.

138 Comprehensive Reviews in Food Science and Food Safety r Vol. 12, 2013 
C 2013 Institute of Food Technologists®
Critical review in fouling and cleaning . . .

An online electrical resistance method has been developed at


the Univ. of Auckland to monitor fouling buildup and removal. A 1.5
Exp. 4.14.2 - 50°C (Re 2340)
Using the fouling surface as an electrode and a reference electrode
on the opposing wall of the test section, the difference in electrical Exp. 4.23.1 - 70°C (Re 3160)
resistance (RE ) across the flow channel can be monitored online

R d (m K/kW)
1.0
(Chen and others 2004). During milk fouling, electrical resistance
increased, similar to thermal resistance (Rf ) measurement. Dur-

2
ing cleaning RE decreased rapidly before any changes in Rf were
detected. The authors determined that NaOH chemical penetra- 0.5
tion into the deposit was not the rate-limiting step in cleaning.
Winquist and others (2005) described the use on a voltammetric
“electronic tongue” that consisted of a series of electrodes im- 0.0
planted in a surface. A measurement series is based on successive -100 400 900 1400 1900 2400 2900 3400 3900 4400
voltage pulses of gradually changing amplitude between which
the base potential is applied, and the current is continuously mea- Time (s)
sured. Electronic tongues were integrated into a dairy process line B Exp. 5.9- 30°C (Re 1600)
Exp. 5.33- 50°C (Re 2340)
and online measurements were taken. A difference in current was Exp. 5.57- 70°C (Re 3160)
2.0
observed between the process stream, water, alkali, and acid, al-
though the investigation back to the clean state was not presented.

Rd (m K/kW)
The authors suggested that the “dirtiness” of the solutions could 1.5
be distinguished by the tongue.

2
1.0
Online surface measurements
Various authors have reported heat transfer measurements dur-
ing cleaning to assess the effect of cleaning parameters on the heat 0.5
transfer coefficient and fouling resistance, Rd , including Gillham
and others (1999, 2000), Christian (2004), and Aziz (2008). An 0.0
example of Rd measured during the cleaning of (a) egg albumin 0 500 1000 1500 2000 2500 3000 3500
and (b) whey protein using 0.5% NaOH at 1.5 L/min is illus-
Time (s)
trated in Figure 14. Rinsing egg albumin at 30 ◦ C did not clean
the surface. Deposit remained on the surface even after 3 h of
Figure 14–Rd profiles for (A) egg albumin gel (from Aziz 2008) and (B)
rinsing. whey protein (from Christian 2004) with different flow temperatures: 30,
Klahre and others (2000) used differential turbidity to moni- 50, and 70 ◦ C using 0.5 wt% NaOH and a flow rate of 1.5 L/min.
tor biofouling on the pipe walls of water systems. The technique
could be considered to study biofouling removal during clean-
ing. The use of a Mechatronic Surface Sensor (MSS) was being
tested to monitor milk components such as calcium phosphate niques discussed that have been investigated in an industrial setting
and whey proteins (Pereira and others 2006) and shampoo (Pereira include:
and others 2009). The sensor measures changes in the vibration
properties of surfaces due to the buildup or removal of fouling r Differential turbidimetry in a paper mill (Klahre and others
layers. 2000),
r Light scattering detection (using a fiber optic probe) in a brewery
Measuring microbial cleanliness (Tamachkiarow and Flemming 2003),
Microbiological enumeration techniques tend to be offline ret- r Biofilm respiration measurement in a bed reactor (Carrión and
rospective techniques. Rinse water samples and surface swabs are others 2003).
plated on selective agar and viable microbes will present themselves
within 3 to 7 d, depending on the organism. Most informative Goode and others (2010) were able to monitor yeast cleaning
swab and plate methods include contact agar method where from stainless steel coupons using a heat flux sensor and the image
the agar plate is pressed directly onto the surface and microbes analysis technique. Less commercially applied techniques discussed
enumerated directly (Salo and others 2008). The cleanliness of a by Janknecht and Melo (2003) include measuring radiation signals
surface can be verified more quickly using ATP bioluminescence (spectroscopy, fluorometry, and photoacoustic spectroscopy) and
that indicates the presence (or absence) of microbes that are alive. electric and mechanical (vibration) signals.
However, the measure of ATP does not indicate microbe speci- The critical part of optimizing a cleaning process would be the
ficity. Microbes have been identified on surfaces using infrared incorporation of cleaning measurements into the process schedule.
spectroscopy (Fornalik 2008) and Raman spectroscopy (Rösch At present, it is usual that cleaning times are set automatically and
and others 2003). Visual assessment, image analysis, and mass are not changed in operation. If measurements made at the start
can all be used to determine the amount of deposit removed (or of cleaning could be used to determine the end point of cleaning
remaining) on a surface after cleaning; however, these methods are with high confidence, it would be possible to develop some form
offline. of a process control strategy that could minimize the cleaning
Janknecht and Melo (2003) published a review discussing on- cost (Yang and others 2008). This approach can be successful if
line biofilm monitoring measurement techniques, known detec- measurements are robust, inexpensive, and taken at the dirtiest
tion limits, and applicability to practical situations. The tech- point within a system.


C 2013 Institute of Food Technologists® Vol. 12, 2013 r Comprehensive Reviews in Food Science and Food Safety 139
Critical review in fouling and cleaning . . .

Summary and Conclusion Any probe use to monitor cleaning needs to be robust and of low
Published information describing the adhesion of fouling ma- cost because cleaning cost is believed to be a relatively low cost of
terials and microorganisms to a range of surfaces used throughout the total cost of production. The robustness and applicability of
the food and beverage industries has been presented and discussed. such novel measurement systems again need to be determined in
Cleaning operations such as CIP are ubiquitously applied to re- an industrial setting.
move unwanted fouling layers in a processing plant to maintain
product safety and process efficiency. However, cost benefit analy-
sis of CIP is not often done; as a result, the route of optimization is
unclear. The grouping of deposits into 3 types has enabled a clear References
presentation of recent studies that have investigated the effect of Akhtar N, Bowen J, Asteriadou K, Zhang Z, Fryer PJ. 2010. Matching the
nano- to the meso-scale: measuring deposit–surface interactions with atomic
CIP parameters. This has enabled parallels in the literature to be force microscopy and micromanipulation. Food Bioprod Process 88:341–8.
drawn. Akhtar N, Bowen J, Robbins PT, Fryer PJ, Asteriadou K, Goode KR. 2012.
The effect of temperature on adhesion forces between surfaces and model
(i) For type 1 deposits, cleaning time seems to be related to Re. foods containing whey protein and sugar. J Food Eng. submitted.
An increase in Re seems to decrease cleaning time according Allie Z, Jacobs EP, Maartens A, Swart P. 2003. Enzymatic cleaning of
to the power law model. It was also seen that increasing the ultrafiltration membranes fouled by abattoir effluent. J Membr Sci
flow rate or wall shear stress and cleaning temperature to 218:107–16.
a mid-range temperature (up to 50 ◦ C) decreases cleaning Asteriadou K, Hasting APM, Bird MR, Melrose J. 2006. Computational
fluid dynamics for the prediction of temperature profiles and hygienic
time. design in the food industry. Food Bioprod Process 84:157–63.
(ii) For type 2 deposits, water rinsing parameters, temperature, Augustin W, Fuchs T, Föste H, Schöler M, Majschak J, Scholl S. 2010.
and wall shear stress seemed to have varied effects on re- Pulsed flow for enhanced cleaning in food processing. Food Bioprod Process
moval. Removal behavior seemed to be dependent on the 88:384–91.
microbial aging time on the surface. NaOH solution re- Aziz NS. 2008. Factors that affect cleaning process efficiency. [PhD thesis].
School of Chemical Engineering, University of Birmingham.
moved type 2 deposits in flowing systems. When consider-
Baier RE. 1980. Substrate influences on adhesion of microorganisms and
ing one chemical concentration, flow and temperature were their resultant new surface properties. In: Bitton G, Marshall KC, editors.
seen to have the biggest effect at the start of cleaning, but Adsorption of microorganisms to surfaces. New York: John Wiley. p.
it was clear that contact time was an important factor as 59–104.
cleaning progressed. Bansal B, Chen XD. 2006. A critical review of milk fouling in heat
(iii) For type 3 deposits, specifically protein, an optimum exchangers. Compr Rev Food Sci Food Safety 5:27–33.
NaOH concentration has been found to occur in numer- Bayod E, Willers EP, Tornberg E. 2008. Rheological and structural
characterization of tomato paste and its influence on the quality of ketchup.
ous studies where excessive chemical material causes forma- LWT Food Sci Technol 41:1289–300.
tion of a deposit difficult to remove. However, increasing Beech IB, Smith JR, Steele AA, Penegar I, Campbell SA. 2002. The use of
wall shear stress and temperature were most beneficial to atomic force microscopy for studying interactions of bacterial biofilms with
cleaning. surfaces. Colloids Surf B Biointerfaces 23:231–47.
Bénézech T. 2001. A method for assessing the bacterial retention ability of
hydrophobic membrane filters. Trends Food Sci Technol 12:36–8.
The findings suggest that optimizing the flow characteristics at a Bénézech T, Lelièvre C, Membré JM, Viet AF, Faille C. 2002. A new test
given temperature and concentration is crucial to achieving fast method for in-place cleanability of food processing equipment. J Food Eng
cleaning in all soil cases. This parameter should be optimized 54:7–15.
in CIP before temperature and or chemical concentration is in- Bird MR. 1992. Cleaning of food process plant. [PhD thesis]. UK:
creased. University of Cambridge.
Novel surface coatings for stainless steel and alternative chemi- Bird MR, Bartlett M. 2002. Measuring and modelling flux recovery during
the chemical cleaning of MF membranes for the processing of whey protein
cals for cleaning are being actively researched at an academic scale. concentrate. J Food Eng 53:143–52.
However, application of these findings has not been adopted in Bird MR, Fryer PJ. 1991. An experimental study of the cleaning of surfaces
industry. The factors affecting the application of research in in- fouled by whey proteins. Food Bioprod Process 69:13–21.
dustry include cost, maintenance, product safety, product quality, Blanpain-Avet P, Migdal JF, Bénézech T. 2009. Chemical cleaning of a
and process reliability. The longevity of surface coatings and the tubular ceramic microfiltration membrane fouled with a whey protein
concentrate suspension—characterization of hydraulic and chemical
traceability of enzymes out of a test system have not been fully cleanliness. J Membr Sci 337:153–74.
demonstrated. However, research in this area is so extensive that Bode K, Hooper RJ, Augustin W, Paterson WR, Wilson DI, Scholl S. 2007.
it is probably only a matter of time before effective solutions are Pulsed flow cleaning of whey protein fouling layers. Heat Transf Eng
found. The pulsing of flow to achieve higher wall shear stress 28:202–9.
within a system looks like a promising route to improve clean- Bott TR. 1990. Fouling notebook. Rugby, UK: Institution of Chemical
Engineers.
ability of process lines. This could be achieved at very low cost
Bott TR. 1995. Fouling of heat exchangers. New York: Elsevier.
because the required equipment is already used in CIP. However,
Bowen WR, Lovitt RW, Wright CJ. 2001. Atomic force microscopy study
longevity of pumps as a result of pulse cleaning has not been fully of the adhesion of Saccharomyces cerevisiae. J Colloid Interface Sci 237:
determined. In the future, minimizing the water load and envi- 54–61.
ronmental impact of cleaning will only become more important. Bremer PJ, Fillery S, McQuillan AJ. 2006. Laboratory scale clean-in-place
Current issues surrounding novel approaches to cleaning will need (CIP) studies on the effectiveness of different caustic and acid wash steps on
to be overcome for application in industrial CIP that will become the removal of dairy biofilms. Int J Food Microbiol 106:254–62.
more important in the future as water becomes less available and/or Burton H. 1967. Deposits from whole milk in heat treatment plant: a review
and discussion. J Dairy Res 34:137–43.
more expensive.
Busscher HJ, Norde W, Sharma PK, van der Mei HC. 2010. Interfacial
There are a number of methods at various stages of commercial- re-arrangement in initial microbial adhesion to surfaces. Curr Opin Colloid
ization for monitoring bulk cleaning and cleaning at the surface. Interface Sci 15:510–7.

140 Comprehensive Reviews in Food Science and Food Safety r Vol. 12, 2013 
C 2013 Institute of Food Technologists®
Critical review in fouling and cleaning . . .

Carrión M, Asaff A, Thalasso F. 2003. Respiration rate measurement in a Fisher RC, Rice FE. 1924. Sweetened condensed milk II. A comparative
submerged fixed bed reactor. Water Sci Technol 47:201–4. study of methods for determining total solids. J Dairy Sci 7:497–502.
Changani SD, Belmar-Beiny MT, Fryer PJ. 1997. Engineering and chemical Fornalik M. 2008. Detecting biofouling in food processing systems.
factors associated with fouling and cleaning in milk processing. Exp Therm Photonics Food 42:58–61.
Fluid Sci 14:392–406. Fryer PJ, Asteriadou K. 2009. A prototype cleaning map: a classification of
Characklis WG. 1980. Biofilm development and destruction. Final Report industrial cleaning processes. Trends Food Sci Technol 20:225–62.
EPRI CS-1554, project RP902-1, Palo Alto, Calif.: Electric Power Fryer PJ, Christian GK, Liu W. 2006. How hygiene happens: physics and
Research Institute. chemistry of cleaning. Int J Dairy Technol 59:76–84.
Chen MJ, Zhang Z, Bott TR. 1998. Direct measurement of the adhesive Gallardo-Moreno AM, González-Martına ML, Bruque JM, Pérez-Giraldo C.
strength of biofilms in pipes by micromanipulation. Biotechnol Tech 2004. The adhesion strength of Candida parapsilosis to glass and silicone as a
12:875–80. function of hydrophobicity, roughness and cell morphology. Colloids Surf A
Chen M-Y, Chen M-J, Lee P-F, Cheng L-H, Huang l-J, Lai C-H, Huang 249:99–103.
K-H. 2010. Towards real-time observation of conditioning film and early Garrett TR, Bhakoo M, Zhang Z. 2008. Bacterial adhesion and biofilms on
biofilm formation under laminar flow conditions using a quartz crystal surfaces. Prog Nat Sci 18:1049–56.
microbalance. Biochem Eng J 53:121–30.
Geesey GG, Gillis RJ, Avci R, Daly D, Hamilton M, Shope P, Harkin G.
Chen XD, Li DXY, Lin SXQ, Özkan N. 2004. On-line fouling/cleaning 1996. The influence of surface features on bacterial colonisation and
detection by measuring electric resistance—equipment development and subsequent substratum chemical changes of 316L stainless steel. Corros Sci
application to milk fouling detection and chemical cleaning monitoring. J 38:73–95.
Food Eng 61:181–9.
Gillham CR. 1997. Enhanced cleaning of surfaces fouled by whey protein.
Chew JYM, Paterson WR, Wilson DI. 2004. Fluid dynamic gauging for [PhD thesis]. U.K.: University of Cambridge.
measuring the strength of soft deposits. J Food Eng 65:175–87.
Gillham CR, Fryer PJ, Hasting APM, Wilson DI. 1999. Cleaning-in-place of
Christian GK. 2004. Cleaning of carbohydrate and dairy protein deposits. whey protein fouling deposits: mechanisms controlling cleaning. Food
[PhD thesis]. U.K.: School of Chemical Engineering, University of Bioprod Process 77:127–38.
Birmingham.
Gillham CR, Fryer PJ, Hasting APM, Wilson DI. 2000. Enhanced cleaning
Christian GK, Changani SD, Fryer PJ. 2002. The effect of adding minerals of whey protein soils using pulsed flows. J Food Eng 46:199–209.
on fouling from whey protein concentrate: development of a model fouling
fluid for a plate heat exchanger. Food Bioprod Process 80:231–9. Goode KR. 2012. Characterising the cleaning behaviour of brewery foulants
to minimise the cost of cleaning in place operations. [EngD thesis].
Christian GK, Fryer PJ. 2006. The effect of pulsing cleaning chemicals on University of Birmingham.
the cleaning of whey protein deposits. Food Bioprod Process 84:
320–8. Goode KR, Asteriadou K, Fryer PJ, Picksley M, Robbins PT. 2010.
Characterising the cleaning mechanisms of yeast and the implications for
Clauberg B, Marcinak M. 2009. Throwing light on photometry. Brewer cleaning in place (CIP). Food Bioprod Process 88:365–74.
Distiller Int 5:48–50.
Gordon PW, Brooker ADM, Chew YMJ, Wilson DI, York DW. 2010.
Cluett JD. 2001. Cleanability of certain stainless steel surface finishes in the Studies into the swelling of gelatine films using a scanning fluid dynamic
brewing process. [Mphil thesis]. South Africa: Faculty of Mechanical gauge. Food Bioprod Process 88:357–64.
Engineering, Rand Afrikaans University.
Grasshoff A. 1997, Cleaning of heat treatment equipment. IDF Monograph,
Cole P, Asteriadou K, Robbins PT, Owen EG, Montague GA, Fryer PJ. Fouling and Cleaning in Heat Exchangers, Brussels, Belgium: International
2010. Comparison of cleaning of toothpaste from surfaces and pilot scale Dairy Federation.
pipe work. Food Bioprod Process 88:392–400.
Grasshoff A. 2002. Enzymatic cleaning of milk pasteurisers. Food Bioprod
Cui ZF, Muralidhara HS. 2010. Membrane technology: a practical guide to Process 80:247–52.
membrane technology and applications in food and bioprocessing. Chapter
10. Oxford, U.K.: Elsevier Science. Güell C, Czekaj P, Davis RH. 1999. Microfiltration of protein mixtures and
the effects of yeast on membrane fouling. J Membr Sci 155:113–22.
De Jong P, van der Linden HJLJ. 1992. Design and operation of reactors in
the dairy industry. Chem Eng Sci 47:3761–8. Guillemot G, Vaca-Medina G, Martin-Yken H, Vernhet A, Schmitz P,
Mercier-Bonin H. 2006. Shear-flow induced detachment of Saccharomyces
Demilly M, Bréchet Y, Bruckert F, Boulangé L. 2006. Kinetics of yeast cerevisiae from stainless steel: influence of yeast and solid surface properties.
detachment from controlled stainless steel surfaces. Colloids Surf B Colloids Surf B 49:126–35.
Biointerfaces 51:71–9.
Guo B, Song S, Chacko J, Ghalambor A. 2005. Offshore pipelines. Chapter
Detry JG, Deroanne C, Sindic M, Jensen BBB. 2009. Laminar flow in radial 16: pigging operations. Oxford, UK: Elsevier.
flow cell with small aspect ratios: numerical and experimental study. Chem
Eng Sci 64:31–42. Henningsson M, Regner M, Östergren K, Trägårdh T, Dejmek P. 2007.
CFD simulation and ERT visualization of the displacement of yoghurt by
Detry JG, Rouxhet PG, Boulangé-Petermann L, Deroanne M, Sindic M. water, on industrial scale. J Food Eng 80:166–75.
2007. Cleanability assessment of model solid surfaces with a radial-flow cell.
Colloids Surf A Physicochem Eng Aspects 302:540–8. Hilbert LR, Bagge-Ravn D, Kold J, Gram L. 2003. Influence of surface
roughness of stainless steel on microbial adhesion. Int Biodeterior Biodegrad
Detry JG, Sindic M, Deroanne C. 2010. Hygeine and cleanability: a focus on 52:175–85.
surfaces. Crit Rev Food Sci Nutr 50:583–604.
Hooper RJ, Liu W, Fryer PJ, Paterson WR, Wilson DI, Zhang Z. 2006.
Dhadwar SS, Bemman T, Anderson WA, Chen P. 2003. Yeast cell adhesion Comparative studies of fluid dynamic gauging and a micromanipulation
on oligopeptide modified surfaces. Biotechnol Adv 21:395–406. probe for strength measurements. Food Bioprod Process 84:
Dror-Ehre A, Adin A, Markovich G, Mamane H. 2010. Control of biofilm 353–8.
formation in water using molecularly capped silver nanoparticles. Water Res Hughes D, Field RW. 2006. Crossflow filtration of washed and unwashed
44:2601–9. yeast suspensions at constant shear under nominally sub-critical conditions. J
EHEDG. 1992. A method for assessing the in-place cleanability of food Membr Sci 280:89–98.
processing equipment. Trends Food Sci Technol 3:325–8. Jackson AT, Low WM. 1982. Circulation cleaning of a plate heat exchanger
EHEDG Yearbook. 2007. Materials of construction for equipment in contact fouled by tomato juice: III. The effect of fluid flow rate on cleaning
with food Trends Food Sci Technol. 18:S40–S50. efficiency. J Food Technol 17:745–52.
Elofsson C, Dejmek P, Paulsson M, Burling H. 1997. Atomic force Janknecht P, Melo LF. 2003. Online biofilm monitoring. Rev Environ Sci
microscopy studies on whey proteins. Int Dairy J 7:813–9. Bio/Technol 2:269–83.
Epstein N. 1983. Thinking about heat transfer fouling: a 5 × 5 matrix. Heat Jefferson K. 2004. What drives bacteria to form a biofilm? FEMS Microbiol
Transf Eng 4:43–56. Lett 236:163–73.
Everaert EPJM, van der Mei HC, Busscher HJ. 1998. Adhesion of yeasts and Jennings WG, Mckillop AA, Luick JR. 1957. Circulation cleaning. J Dairy
bacteria to fluoro- alkylsiloxane layers chemisorbed on silicone rubber. Sci 40:1471–9.
Colloids Surf B Biointerfaces 10:179–90. Jensen BBB, Friis A. 2005. Predicting the cleanability of mix-proof valves by
Fickak A, Al-Raisi A, Chen WD. 2011. Effect of whey protein use of wall shear stress. J Food Process Eng 28:89–106.
concentration on the fouling and cleaning of a heat transfer surface. J Food Jensen BBB, Friis A. 2004. Critical wall shear stress for the EHEDG test
Eng 104:323–31. method. Chem Eng Process 43:831–40.


C 2013 Institute of Food Technologists® Vol. 12, 2013 r Comprehensive Reviews in Food Science and Food Safety 141
Critical review in fouling and cleaning . . .

Jensen BBB, Stenby M, Nielsen DF. 2007. Improving the cleaning effect by Quain D, Storgårds E. 2009. The extraordinary world of biofilms. Brewer
changing average velocity. Trends Food Sci Technol 18:S52–S63. Distiller Int 5:31–33.
Klahre J, Flemming M, Flemming H-C. 2000. Monitoring of biofouling in Quarini J. 2002. Ice-pigging to reduce and remove fouling and to achieve
papermill process waters. Water Res 34:3657–65. clean-in-place. Appl Thermal Eng 22:747–53.
Kim JY, Lee JH, Chang IS, Lee JH, Yi CW. 2011. High voltage impulse Rao MA. 1999. Rheology of fluid and semisolid foods principles and
electric fields: disinfection kinetics and its effect on membrane bio-fouling. applications. U.S.A.: Aspen Publishers.
Desalination 283:111–6. Reynolds TB, Fink GR. 2001. Bakers’ yeast, a model for fungal biofilm
Le Gentil C, Sylla Y, Faille C. 2010. Bacterial re-contamination of surfaces formation. Science 291:878–81.
of food processing lines during cleaning in place procedures. J Food Eng Robbins PT, Elliott BL, Fryer PJ, Belmar MT, Hasting APM. 1999. A
96:37–42. comparison of milk and whey fouling in a pilot scale heat exchanger:
Lelieveld HLM, Mostert MS, Holah J. 2005. Handbook of hygiene control implications for modelling and mechanistic studies. Food Bioprod Process
in the food industry. EHEDG, 192 – 208. Cambridge, U.K.: Woodhead. 77:97–106.
Lelièvre C, Antonini G, Faille C, Bénézech T. 2002. Cleaning-in-place Rösch P, Schmitt M, Keifer W, Popp J. 2003. The identification of
modelling of cleaning kinetics of pipes soiled by bacillus spores assuming a microorganisms by micro-Raman spectroscopy. J Mol Struct
process combining removal and deposition. Food Bioprod Process 661–662:363–9.
80:305–11. Rosmaninho R, Santos O, Nylander T, Paulsson M, Beuf M, Benezech T,
Liu W, Aziz NA, Zhang Z, Fryer PJ. 2007. Quantification of the cleaning of Yiantsios S, Andritsos N, Karabelas A, Rizzo G, Müller-Steinhage H, Melo
egg albumin deposits using micromanipulation and direct observation LF. 2007. Modified stainless steel surfaces targeted to reduce fouling –
techniques. J Food Eng 78:217–24. evaluation of fouling by milk components. J Food Eng 80:
Liu W, Christian GK, Zhang Z, Fryer PJ. 2002. Development and use of a 1176–87.
micromanipulation technique for measuring the force required to disrupt Saikhwan P, Mercade-Prieto R, Chew YMJ, Gunasekaran S, Paterson WR,
and remove fouling deposits. Food Bioprod Process 80:286–91. Wilson DI. 2010. Swelling and dissolution in cleaning of whey protein gels.
Liu W, Fryer PJ, Zhang Z, Zhao Q, Liu Y. 2006. Identification of cohesive Food Bioprod Process 88:375–83.
and adhesive effects in the cleaning of food fouling deposits. Innov Food Sci Salo S, Friis A, Wirtanen G. 2008. Cleaning validation of fermentation tanks.
Emerg Technol 7:263–9. Food Bioprod Process 86:204–10.
Liu M, Li X, Lin R, Nie W, Zhang N, Ling N. 2004. Fouling prevention Santos O, Nylander T, Rosmaninho R, Rizzo G, Yiantsios S, Andritsos N,
with fluidised particles in evaporation of traditional Chinese medicine Karabelas A, Müller-Steinhagen H, Melo L, Boulangé-Petermann L, Gabet
extract. China Particuol 2:81–3. C, Braem A, Trägårdh C, Paulsson M. 2004. Modified stainless steel surfaces
Lorite GS, Rodrigues CM, de Souza AA, Kranz C, Mizaikoff B, Cotta MA. targeted to reduce fouling—surface characterization. J Food Eng 64:
2011. The role of conditioning film formation and surface chemical changes 63–79.
on Xylella fastidiosa adhesion and biofilm evolution. J Colloid Interface Sci Schlüsser, HJ. 1976. Zur Kinetik von Reinigungsvorgängen an festen
359:289–95. Oberächen. Brauwissenschaft 29:263–8.
Mercier-Bonin M, Ouazzani K, Schmitz P, Lorthois S. 2004. Study of Scottish & Newcastle Breweries. 2008. CIP Philosophy, Scottish &
bioadhesion on a flat plate with a yeast/glass model system. J Colloid Newcastle Achiever Database.
Interface Sci 271:342–50. Sharma A, Garg D, Gupta JP. 1982. Solidification fouling of paraffin wax
Mores WD, Davis RH. 2002. Yeast foulant removal by backpulses in from hydrocarbons. Lett Heat Mass Transf 9:209–19.
crossflow microfiltration. Journal Membr Sci 208:389–404. Shorrock CJ, Bird MR. 1998. Membrane cleaning: chemically enhanced
Morison KR, Thorpe RJ. 2002. Liquid distribution from cleaning-in-place removal of deposits formed during cell harvesting. Food Bioprod Process
sprayballs. Food Bioprod Process 80:270–5. 76:30–8.
Mozes M, Marchal F, Hermesse MP, van Haecht JL, Reuliaux L, Leonard Simões M, Pereira MO, Vieira MJ. 2005. Effect of mechanical stress on
AJ, Rouxhet PG. 1987. Immobilization of microorganisms by adhesion: biofilms challenged by different chemicals. Water Res 39:5142–52.
interplay of electrostatic and nonelectrostatic interactions. Biotechnol Stewart JC, Seiberling DA. 1996. Cleaning in place. Chem Eng 103:72–79.
Bioeng 30:439–50.
Tamine AV. 2008. Cleaning-in-place: dairy, food and beverage operations.
Müller-Steinhagen H. 2000. Heat exchanger fouling mitigation and cleaning Society of Dairy Technology Series. London: Wiley-Blackwell.
technologies, 1–28. Rugby, U.K.: Institution of Chemical Engineers.
Tamachkiarow A, Flemming HC. 2003. On-line monitoring of biofilm
Orgaz B, Kives J, Pedregosa AM, Monistrol IF, Laborda F, SanJosé C. 2006. formation in a brewery water pipeline system with a fibre optical device.
Bacterial biofilm removal using fungal enzymes. Enzyme Microb Technol Water Sci Technol 47:19–24.
40:51–6.
Timperley AW, Boution F, Bénézech T, Carpentier B, Curiel GJ, Haugan
Othman AM, Asteriadou K, Robbins PT, Fryer PJ. 2010. Cleaning of sweet K, Hofman J. 2000. A method for the assessment of in-place cleanability of
condensed milk deposits on a stainless steel surface. In: Wilson DI, Chew food processing equipment. EHEDG. 2nd ed. Chipping, Campden,
YMJ, editors. Proceedings of the Fouling & Cleaning in Food Processing England: CCFRA Technology Ltd. p. 1–14.
Conference. Cambridge: Department of Chemical Engineering. p. 174–82.
Timperley DA, Smeulders CNM. 1988. Cleaning of dairy HTST plate heat
Palabiyik I, Olunloyo B, Fryer PJ, Robbins PT. 2012. Flow regimes in the exchangers: optimisation of the single-stage procedure. J Soc Dairy Technol
emptying of pipes filled with viscoelastic material. Int J Multiphase Flow. 41:1–7.
submitted.
Tse KL, Pritchard AM, Fryer PJ. 2003. The rate and extent of fouling in a
Parbhu A, Hendy S, Danne M. 2006. Reducing milk protein adhesion rates. single-tube wort boiling system. Food Bioprod Process 81:13–22.
A transient surface treatment of stainless steel. Food Bioprod Process
84:274–8. Tuladhar TR. 2001. Development of a novel sensor for cleaning studies.
[PhD thesis]. U.K.: University of Cambridge.
Pereira A, Mendes J, Melo LF. 2009. Monitoring cleaning-in-place of
shampoo films using nanovibration technology. Sens Actuators B Tuladhar TR, Paterson WR, Wilson DI. 2002. Investigation of alkaline
136:376–82. cleaning-in-place of whey protein deposits using dynamic gauging. Food
Bioprod Process 80:199–214.
Pereira A, Rosmaninho R, Mendes J, Melo LF. 2006. Monitoring deposit
build-up using a novel mechatronic surface sensor (MSS). Food Bioprod Yoo JI, Chen XD, Mercadé-Prieto R, Wilson DI. 2007. Dissolving
Process 84:366–70. heat-induced protein gel cubes in alkaline solutions under natural and
forced convection conditions. J Food Eng 79:1315–21.
Pereni CI, Zhao Q, Liu Y, Abel E. 2006. Surface free energy effect on
bacterial retention. Colloids Surf B Biointerfaces 48:143–7. Van Asselt AJ, Van Houwelingen G, Te Giffel MC. 2002. Monitoring system
for improving cleaning efficiency of cleaning-in-place processes in dairy
Petrus HB, Chen HLV, Norazman N. 2008. Enzymatic cleaning of environments. Food Bioprod Process 80:276–80.
ultrafiltration membranes fouled by protein mixture solutions. J Membr Sci
325:783–92. Vinogradov AV, Winston M, Rupp CJ, Stoodley P. 2004. Rheology of
biofilms formed from the dental plaque pathogen S. Mutans. Biofilms
Plett EA. 1985. Cleaning of fouled surfaces. In: Lund DB, Plett EA, Sandu 1:49–56.
C, editors. Fouling and cleaning in food processing. Wisconsin: University
of Madison. Whitehead KA, Rogers D, Colligon J, Wright C, Verran J. 2006. Use of the
atomic force microscope to determine the effect of substratum surface

142 Comprehensive Reviews in Food Science and Food Safety r Vol. 12, 2013 
C 2013 Institute of Food Technologists®
Critical review in fouling and cleaning . . .

topography on the ease of bacterial removal. Colloids Surf B Biointerfaces Zhao Q, Liu Y. 2006. Modification of stainless steel surfaces by electroless
51:44–53. Ni-P and small amount of PTFE to minimize bacterial adhesion. J Food
Whitehead KA, Verran J. 2006. The effect of surface topography on the Eng 72:266–72.
retention of microorganisms. Food Bioprod Process 84:253–9. Zhao Q, Liu Y, Wang C. 2005a. Development and evaluation of electroless
Wilson DI, Le BL, Dao HAD, Lai KY, Morison KR, Davidson JF. 2012. Ag-PTFE composite coatings with anti-microbial and anti-corrosion
Surface flow and drainage films created by horizontal impinging liquid jets. properties. Appl Surf Sci 252:1620–7.
Chem Eng Sci 68:449–60. Zhao Q, Liu Y, Wang C, Wang S, Müller-Steinhagen H. 2005b. Effect of
Winquist F, Bjorklund R, Krantz-Rülcker C, Lundströma C, Östergren K, surface free energy on the adhesion of biofouling and crystalline fouling.
Skoglund T. 2005. An electronic tongue in the dairy industry. Sens Chem Eng Sci 60:4858–65.
Actuators B 111–112:299–304. Zhao Q, Wang C, Liu Y, Wang S. 2007. Bacterial adhesion on the
Xiaokai X, Chongfang M, Yongchang C. 2005. Investigation of the metal-polymer composite coatings. Int J Adhes Adhes 27:
electromagnetic antifouling technology for scale prevention. Chem Eng 85–91.
Technol 28:1540–5. Zhao Q, Wang S, Müller-Steinhagen H. 2004. Tailored surface free energy
Yang A, Martin EB, Montague GA, Fryer PJ. 2008. Towards improved of membrane diffusers to minimize microbial adhesion. Appl Surf Sci
cleaning of FMCG plants: a model-based approach. Comput Aided Chem 230:371–8.
Eng 25:1161–6.


C 2013 Institute of Food Technologists® Vol. 12, 2013 r Comprehensive Reviews in Food Science and Food Safety 143

You might also like