Anda di halaman 1dari 23

Fouling and Cleaning Studies in the Food and Beverage Industry Classied 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 signicant industrial problem, and as the cost of water and chemical disposal increases, the problem is becoming more signicant. Current literature on water-based cleaning is reviewed here according to the classication 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 benecial. (ii) For type 2 deposits (biolms)Removal behavior of biolms 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 ow and temperature. (iii) For type 3 deposits (that require a cleaning chemical)For specically, protein-based systems excessive chemical forms a deposit difcult to remove. Increasing wall shear stress and temperature was most benecial to cleaning rather than concentration. The action of temperature can reduce the use of a chemical for type 2 and type 3 soils. The ndings suggest that the right combination of ow 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 1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 2. Fouling studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 2.1 Adhesion of microorganisms to surfaces . . . . . . . . . . . . . . . . 4 2.3 Preventing fouling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 2.3.1 Process surface modication . . . . . . . . . . . . . . . . . . . . . . 5 2.3.2 Process alterations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6 3. Cleaning. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .6 3.1 Product recovery . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 3.2 The effect of CIP parameters on type 1 removal . . . . . . . 10 3.2.1 Flow and wall shear stress . . . . . . . . . . . . . . . . . . . . . . . . 11 3.2.2 Temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 3.2.3 Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 3.3 The effect of CIP parameters on type 2 and type deposit removal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 3.3.1 Membrane cleaning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 3.3.2 Water-rinsing of hard surfaces . . . . . . . . . . . . . . . . . . . . 14

3.3.3 Chemical effects on the cleaning of type 2 deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 3.3.4 Chemical effects on the cleaning of type 3 deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 4. Novel cleaning approaches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 4.1 Increasing boundary layer disruption. . . . . . . . . . . . . . . . . .16 4.2 Alternative cleaners to reduce environmental impact . . . 16 4.3. Other studies related to cleaning behavior . . . . . . . . . . . . 17 4.3.1 Deposit shear . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 4.3.2 Deposit deformation and strength . . . . . . . . . . . . . . . . 17 5. Measuring cleaning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 5.1 Online bulk measurements . . . . . . . . . . . . . . . . . . . . . . . . . . 18 5.2 Online surface measurements . . . . . . . . . . . . . . . . . . . . . . . . 19 5.3 Measuring microbial cleanliness . . . . . . . . . . . . . . . . . . . . . . 19 6. Summary and conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

Introduction

The success of a branded fast moving consumer goods (FMCGs) business depends fundamentally on product quality and safety conMS 20120483 Submitted 3/29/2012, Accepted 10/14/2012. Authors are with formance at a required level. Poor cleaning or hygiene conforSchool 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 . . .

Scienc invesgaon

Invesgate design Invesgate process parameters

Dene CIP standard Best process design Best process parameters

Establish standard at all sites

Input from science

Input from pracce

NON CONFORMANCE e.g. Poor maintenance, control or training Why not?

By site educaon, training and empowerment and recording site CIP condions

NO

Is hygiene being achieved?

YES

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

should be fundamental research and development obtained from both practice and science that are integrated and applied in-plant to provide the optimum cleaning protocol. In food and beverage manufacturing operations, cleaning-in-place (CIP) is used to remove residual product, fouling, and microbes that remain in the process line from production. The act of cleaning therefore maintains product quality, safety, and production efciency. During CIP, water and/or chemical solution is circulated around plant process equipment. With large-scale manufacturers, the process is generally fully automated. A typical CIP philosophy in industry is that of Scottish & Newcastle Breweries (2008):

ensure all production, processing, and packaging plant is cleaned by a standard regime and to a schedule which ensures cleanliness and microbiological integrity at all times; with minimum cost, energy, and delay to production in a manner which ensures human, plant, product, and environmental safety.

identify the best way to clean a processing plant from experiments of different plants. The direct selection of cleaning protocols is not always possible. In practice, cleaning protocols can only be developed semiempirically in industry. In most cases, CIP cannot be optimized in situ because of the risk posed of compromising existing cleanliness. Fryer and Asteriadou (2009) suggest a classication of cleaning problems in terms of cleaning cost and soil complexity. A diagrammatic representation of this relationship is presented in Figure 2. This classication enables the nature of a foulant to be related to the type of cleaning employed, and therefore, the cost. This classication also indicates the environmental impact of the type of cleaning employed; complex soils require chemical and thermal cleaning that lead to a high cleaning cost and high environmental impact. Three deposit types were chosen to represent a broad range of cleaning problems seen in food, beverage, and personal care products manufacturing:

(i) Type 1: Viscoelastic or viscoplastic uids such as yogurt and A signicant body of cleaning knowledge exists within individtoothpaste that can be rinsed from a process surface with ual manufacturers, equipment suppliers, and chemical companies; water. however, the determined cleaning regimens have often been kept (ii) Type 2: Microbial and gel-like lms such as biolms and condential and plant-specic. This has resulted in independent polymers removed in part by water and in part by chemical. development of cleaning operations. Organizations such as the European Hygienic Engineering and Design Group (EHEDG) (iii) Type 3: Solid-like cohesive foulants formed during thermal processing such as milk pasteurization and brewery have produced extensive guidelines on the types of surface and wort evaporation. These operations mostly require chemical equipment that are easy to clean, such as detailed in the EHEDG removal. Yearbook (2007). CIP tends to follow a similar series of steps for a prescribed Yang and others (2008) also classied cleaning optimization time and at a prescribed ow rate, temperature, and chemical methods, here into 2 types of investigation: concentration known to give a repeatable level of cleanliness. It (i) Engineering investigations: reducing energy, time, and cost in is not yet possible to predict before an operation how a given established cleaning operations. piece of equipment could foul and be cleaned. It is difcult to
122 Comprehensive Reviews in Food Science and Food Safety r Vol. 12, 2013
2013 Institute of Food Technologists

Input from pracce

LACK OF FUNDAMENTALS e.g. The eect of CIP parameters on dierent fouling condions

Critical review in fouling and cleaning . . .

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

(ii) Scientic investigations: achieving cleanliness or a cleaning time as a function of inuencing factors; for example, wall shear stress, temperature, surface type, and nish. The aim of this review was to provide an overview of current knowledge on cleaning solutions classied by soil cleaning type. This novel classication is hoped to highlight new CIP optimization opportunities for industry and any future research in the eld. Current knowledge of fouling prevention and novel cleaning methods are also discussed here.

r Growth of biolmoften after the formation of a condir Accumulation of material in stagnant or low-ow areas of r Loss of membrane activity.

tioning layer of protein onto the surface. equipment.

Fouling is a costly problem in the food, beverage, and other industries, which is often unavoidable due to the heat treatment that often has to be given to products to develop certain colors and avors and ensure safety. By denition, foods are sources of nutrients favorable not only to people but also to microbes that stick to process surfacesso microbial adhesion to surfaces and subsequent growth are important pheFouling Studies Fouling is dened as the unwanted buildup of material on a nomena. The economic penalties of fouling in heat exchanguller-Steinhagen (2000) and can be surface. The fouling process generally involves a number of steps ers were discussed by M summarized: (Epstein 1983): (i) (ii) (iii) (iv) (v) surface conditioning, mass transfer of species to the surface, surface deposition, deposit aging, and possible removal. (i) Capital expenditure, due to: (a) Excess heat transfer surface area compensating for the occurrence of fouling. This has been estimated as an average of 30% additional capital cost. (b) Higher transport and installation costs for bigger and heavier equipment. (c) Cleaning systems, including their installation and maintenance costs. (i) Fuel costIf extra energy (such as steam) is required to keep the fouled heat exchanger operating for the required performance. (ii) Maintenance costOf the heat exchanger, cleaning system, and any ancillary equipment in the process (and cleaning) loop, for example, chemical tank level probes, ow meters, interface probes, and boilers.

There is also a classication of fouling mechanisms demonstrated by Bott (1990) detailed in Table 1. Fouling problems that have been reported in the food and beverage industries include (but this is by no means a complete list):
r r r r r

Protein and mineral deposition in heat exchangers. Ice buildup in freezers. Scale buildup in cooling water systems. Fat burn-on in ovens. Product solidication.

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 1Fouling mechanisms: adapted from Bott (1990) and Sharma and others (1982). Fouling mechanism Crystallization Underlying process

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 uid or components of the uid solidify onto the surface, this is called solidication 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 lms consisting of microorganisms and their products, called biolm. Chemical reaction at Reaction of some part of the ow to generate insoluble material. The deposit formed on the surface (particularly heat uid/surface interface transfer surfaces) has a different composition to the process uid (for example, in petroleum rening, polymer production, and dairy plants). Corrosion The material of the heat transfer surface is involved in reactions with components of the uid to form corrosion products on the surface, a specic type of chemical reaction fouling. Freezing Deposit formed from a frozen layer of the process uid, for example, ice from water or solid fats from a food uid.

(iii) Cost due to production lossCost 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 adhesion and biolm formation easier. The sequence of events that as compared to the actual production cost. occur during lm formation is discussed by Busscher and others Accurate measurement of the effects of fouling and the efciency (2010) and Chen and others (2010) who show the kinetics of lm is critical. Changes in heat transfer efciency 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 coefcient, (U0 ), to that at and Fink (2001) proved that Bakers yeast can initiate biolm fortime 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) = of cells on stainless steel and aluminum at pH 3 and pH 5 and 6. U U0 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 (): Bi = Rf .U0 , where Rf = x/ 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 other authors to readily attach to stainless steel, plastics, elastomers 1 1 Rd = (2) (Guillemot and others 2006), and glass (Mercier-Bonin and othUc Ut ers 2004), all of which are used extensively in FMCG industries. where Ut is the heat transfer coefcient at time t and Uc the heat The effect of cleaning parameters on yeast removal from process transfer coefcient of the nal clean system, so the rate of change surfaces is discussed in later sections. of this is a measure of cleaning. Product contact surface nishes with a roughness (Ra ) value The rate and extent of fouling and cleaning is often classied of up to 0.8 m are recommended (Lelieveld and others 2005), in terms of uid ow, either in terms of the Reynolds number which is often called 2B nish of stainless steel. Surface rough(Re = vd/, where and are the density of viscosity of a ness exists in 2 principal planes, one perpendicular to the surface uid owing at mean velocity v through a system of characteristic described as height deviation and one in the plane of the surlength 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 rediscussedto convert to velocity requires knowledge of density tention has been investigated most thoroughly. Hilbert and others and viscosity of the uid, 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 m) on retaining various microbes. The surfaces also had a conditioning layer. The retention of microbes (measured by indiAdhesion 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 nish on the fouling and cleaning of a beer fermenter. Surface total attractive force (Bott 1995). Microbes have a natural afnity nishes investigated included 2B milled stainless steel and meto 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 B en ezech 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 biolms (adhesive and cohesive all nishes, one quarter of the vessel from top to bottom reprecommunities of microbes) that become difcult to remove from sented by each surface nish. 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 biolms in industry, fouling mechanisms and the level of deposition was heavy. He also found that all the surmethods 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

2013 Institute of Food Technologists

Critical review in fouling and cleaning . . .


(prerinse, caustic, water, acid, water, and sanitizer). However, the number of viable microbes was found to decrease in the cone at the bottom of the vessel. Gallardo-Moreno and others (2004) investigated the effect of surface roughness by comparing yeast adhesion to glass (Ra 0.8 m and hydrophilic) and silicone rubber (SR) (Ra 0.61 m and hydrophobic). The authors found larger adhesion rates for SR, and at 37 C rather than 22 C. Whitehead and others (2006) investigated Pseudomonas aeruginosa (rods of 1 m width and 3 m length) and Staphylococcus aureus (1 m sphere) retention on a titanium dioxide surface: smooth with dened surface features (pits) of 0.5 m. S. aureus cells were removed more easily from the smooth surface, whereas P. aeruginosa cells were removed more easily from the defects. Whitehead and Verran (2006) also reviewed the effect of Ra and topography on microbial retention. Research suggests that surfaces with a Ra value close to the cell size see increased retention on the surface. For example, yeasts were found to require larger defects (5 m) for retention and smaller daughter cells were retained in smaller defects (2 m). Rod-shaped cells seemed to orient themselves in grains and grooves of similar size. yeast adhesion. Overall surface energy was 50 to 60 and 25 mJ/m for cell adhesion on plastic and glass, respectively. Coating both surfaces changed the free energy of the system resulting in a decrease to 35 to 40 mJ/m for plastic and an increase to 30 to 40 mJ/m for glass. Yeast adhesion was signicantly reduced on the plastic surface coated in the peptide and increased on the glass surface. Changes in surface roughness and hydrophobicity due to the coating will also have contributed to adhesion. Quain and Storg ards (2009) mentioned the testing of functional materials in the lab and in brewery dispense lines such as hydrophobic uoropolymer coatings, photocatalytic titanium dioxide coatings, and the inclusion of antimicrobial silver ions (0.042%) in stainless steel. The latter was shown to reduce the number of adhering bacteria by 99% compared to normal stainless steel. However, the effect decreased with time. The inuence of surface energy on adhesion is well known in marine and medical biofouling and is characterized by the Baier curve (Baier 1980). This curve demonstrates the weakest adhesive strength of bacteria to be at surface energies of around 25 mN/m. Equations dening possible minimum adhesion energies between a deposit and the surface have been developed. The following equation has been derived:
LW S =

Preventing fouling If fouling were not to occur, there would be little need for cleaning. Broadly speaking, 2 methods for preventing fouling have been approached in the literature:
(i) Functional surfacesFor example, smooth surfaces with specic nish, topography, hydrophobicity, or surface charge. Nonstick surfaces are designed to have a specic surface energy to minimize fouling. (ii) Processing alterationsFor example, changing product ow characteristics, holding times, transient times, and other process parameters designed to minimize fouling.
Process surface modication. A hygienic surface needs to be smooth, easy to clean, able to resist wear, and retain its hygienic qualities. Stainless steel is the most common food contact material used in the industry, being stable at a variety of temperatures, inert, relatively resistant to corrosion, and it may be treated mechanically or electrolytically to obtain a range of nishes (Akhtar and others 2010). The wettability of a surface is dependent on its surface energy. A surface with a high surface energy is hydrophilic and a drop of cleaning uid will spread over the surface. A low-energy surface is hydrophobic and a drop of water will not spread. Water partially wets glass and acrylic and does not wet Teon (PTFE) surfacesbut surfactants are often added to commercial cleaning agents to improve wetting. Wetting is determined by the nature of both the liquid and the solid substrate. The cleanability (and disinfectability) of stainless steel has been compared with those of other materials, and is comparable to glass when cleaning microbes, and signicantly better than polymers, aluminum, or copper (Akhtar and others 2010). Microbes are known to readily attach to SR. Everaert and others (1998) absorbed long uorocarbon chains (Ar-SR-C8 F17 ) to SR used in prosthetics in an attempt to reduce the number of adhering microbes. They found that the initial adhesion rate of Streptococcus bacteria to the treated rubber was signicantly reduced, from around 2500 to 900 cm2 s1 , without a conditioning lm of saliva and 400 cm2 s1 with a conditioning lm of saliva. The adhesion rate of Candida species to treated rubber was also reduced compared to untreated rubber. Dhadwar and others (2003) investigated the effect of oligopeptide treatment of glass (hydrophilic) and plastic (hydrophobic) on
2013 Institute of Food Technologists

1 2

LW D +

LW F

(3)

LW LW LW , D , and F are the Lifshitzvan der Waals (LW) where S surface free energy of the surface, deposit, and uid, respectively, and which can be quantied from contact angle measurements (Zhao and others 2004). Liu and others (2006) studied the interactions of 316 L stainless steel with baked and unbaked tomato deposit: a minimum removal energy range of 20 to 25 mN/m was found in both cases. Either side of this surface energy range, the adhesive strength of the deposit on the surface increased. Zhao and others (2005a) found that stainless steel surfaces coated with Ag-PTFE reduced Escherichia coli attachment by 94% to 98%, compared with silver coating, stainless steel, or titanium surfaces. A surface with energy of 24.5 mN/m roughly matching the theoretical minimum adhesion energy of E. coli, 28.3 mN/m, was achieved. Composite coatings using nickel, phosphorus, copper, and PTFE were also used by Zhao and others (2005b) and Zhao and Liu (2006) to create surfaces with specic energies shown to reduce biofouling. A major EU project (MODSTEEL) developed and studied a wide range of surfaces and how they might reduce fouling from milk (see Santos and others 2004 and Rosmaninho and others 2007). Work by Pereni and others (2006) conrmed the effect of surface free energy in minimizing P. aeruginosa adhesion to a range of coatings including silicone, polished and nonpolished stainless steel, PFA (peruoroalkoxy polymer) and PTFE nickel, phosphorus, and aluminum composite coatings. The total surface free energy was in the range 17.2 to 48.3 mN/m, as shown in Figure 3(A). Minimum retention of bacteria was found at 20 to 27 mN/m. Silicone had a surface free energy of around 20 mN/m and the lowest colony forming units (CFUs) count. Surface free energy has been shown to be the parameter dominating E. coli adhesion over a range of metalpolymer coatings, and a minimum adhesion energy of 25 mN/m was found (Zhao and others 2007), as shown in Figure 3(B). Parbhu and others (2006) used a transient treatment to modify a stainless steel surface. The treatment was present during the processing cycle and removed at high pH during alkaline cleaning.

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 signicant biolm buildup. Tse and others (2003) found that in a 2-phase (liquid-vapor) wort boiling system, the wall temperature did not signicantly affect the rate of fouling. Under conditions where vapor was condensed at lower ow velocities (0.07 and 0.14 m/s), the initial fouling phase was more rapid at the higher ow velocity. The authors found that the initial fouling rate was halved as the ow velocity was doubled. These ndings suggest that circulating uid at a fast ow rate would reduce fouling. The authors also found that at the lowest ow 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 multilayered. 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 ow (liquid-vapor) and 3-phase ow (liquid vaporsolid) during the evaporation of Gengnianan extract. The solid phase was added as inert solid particles. The 2-phase ow system generated fouling in 15 h, whereas the 3-phase ow system generated fouling after 60 h. Modifying the process by using electric elds has also been discussed. 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 temperatures and less fouling initially. However, fouling in the bulk is easily transferred to the surface, resulting in fouling (Bansal and Chen 2006). Kim and others (2011) demonstrated that an electric eld could be used to control membrane fouling with E. coli. E. coli cell suspensions were treated by an electric eld prior to ltration. The ux of the suspension was maintained throughout the ltration period due to larger fouling particles reducing cake resistance. Cell death also increased with increasing electric eld strength from 5 to 20 kV/cm. Flux of the untreated E. coli suspension decreased abruptly after the onset of ltration. Xiaokai and others (2005) investigated the effect of electromagnetic treatment of water to minimize scale formation in the tubes of a plate heat exchanger (PHE). The technology is termed electromagnetic antifouling (EAF). The treatment was shown to aggregate particles in the ow that led to reduced precipitation at the wall.

Figure 3(A) P. aeruginosa AK1 retention on investigated surfaces compared with the total surface free energy (adapted from Pereni and others 2006). (B) Effect of surface free energy on E. coli adhesion (adapted from Zhao and others 2007).

The treatment was shown to reduce the interaction potential between stainless steel and phosphate anions resulting in signicant reductions in fouling rates. Akhtar and others (2010) compared adhesion of a range of food and personal care foulants to different surfaces. Particle tips of different materials were attached to an atomic force microscope (AFM) cantilever to study the detachment from toothpaste and some confectionery components: Turkish delight, caramel, and sweetened condensed milk (SCM). The study did reveal signicantly different detachment forces for the same deposit from different surface types (see Figure 4). Caramel and SCM seemed to be more difcult to detach from glass than stainless steel. It was possible to relate data from the AFM to measurements taken on a millimeter scale using micromanipulation probes (Liu and others 2002, 2006, 2007). Akhtar and others (2012) describe further research using AFM to study food adhesion to process surfaces. Process alterations. Dror-Ehre and others (2010) tested the effect of biolm development of P. aeruginosa when pretreated in an aqueous solution of molecularly capped silver nanoparticles (MCNPs). Under specic conditions, cells and surfaces incubated for 39 h at 37 C, Ag-MCNPs retarded biolm formation even when a high percentage of planktonic P. aeruginosa cells survived pretreatment with Ag-MCNPs. At the various incubation times, a stable, low value of biomass was formed that could be easily removed. The authors found, from micrographs of pretreated cells, that the intracellular material was pushed toward the peripheral parts of the

Cleaning
No economically viable fouling prevention method is yet to be demonstrated in industry. Should one of the modied surface methods prove economic, then the problem will be greatly reduced. Understanding the cleanability of surfaces requires combining understanding of surface chemistry and engineering, the deposit and the cleaning uid (for a recent review of cleanability, see Detry and others 2010). Further research discussed here considers the ndings of studies relevant to optimizing cleaning. One signicant issue is determination of the correct cleaning time. A deposit that has aged on a surface is more difcult to remove than fresh material on a surface, so cleaning is encouraged after production. Aging of a particular soil type could make a deposit harder to remove from a process surface. For example, a type 1 soil could become a type 2 soil over time and heating may result in a type 3 soil. Goode and others (2010) found that in
2013 Institute of Food Technologists

126 Comprehensive Reviews in Food Science and Food Safety r Vol. 12, 2013

Critical review in fouling and cleaning . . .

1 Caramel 0.1 F/R(N/m) SCM Turkish delight Toothpaste

Stainless steel PTFE Glass

0.01

0.001

0.0001
Figure 4Force of attraction between stainless steel, PTFE (uorinated 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 classication tool and forms the basis for the structure of this review. Examples cleaned, classied as type A and type B foulants: (i) Type AFormed during fermentation above the beer level of each deposit type include: at the top of the vessel, (ii) Type BResidual yeast attached to the vessel wall and cone below the beer level during emptying. Type B fouling is shown by Salo and others (2008), as seen in Figure 5(A), while an example of type A foulant viewed from a 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. As such, type B foulant can be removed by the falling lm in a tank, whereas type A foulant may require a larger impact force for removal or a combination of water and chemical rinses for complete removal (Goode and others 2010). Similarly, Liu and others (2002) found that the force required to remove a tomato deposit from a surface increased with time until after about 200 min of heating it remained constant. Automated CIP has been widely applied in dairies, food processing, brewing, and wine processing for the last 50 y to return the plant to a clean state (Stewart and Seiberling 1996). During CIP, water and chemicals are circulated around the plant for a prescribed duration (Tamine 2008). The CIP factors found to determine cleaning can be described by Sinners circle, a circle of the cleaning parameters: mechanical action, chemical action, time, and temperature (Lelieveld and others 2005). Cleaning can also be dependent on geometry. In a pipe, the contribution of the cleaning factors is equal. In a pipe dead leg, time determines cleaning (Lelieveld and others 2005). A number of attempts have been made to try to incorporate computational models into the design process, as shown by Asteriadou and others (2006) and Jensen and Friis (2005). This approach will become more important as understanding of the processes in cleaning increases. Rheological characterization of materials enables their classication. Materials within a similar class may have similar cleaning behavior, according to the classication by Fryer and Asteriadou (2009). Vinogradov and others (2004) characterized the rheology of a dental plaque biolm. Biolm rheology has been viscoelastic, temperature-dependent and/or time-dependent (Rao 1999). Characklis (1980) compared the elastic and viscous modules obtained for a biolm and a cross-linked protein gel, brinogen. The elastic modulus was the same order of magnitude for the protein gel and the biolm. The cleaning map, presented in Figure 2 (Fryer
2013 Institute of Food Technologists

(i) Type 1: toothpaste, tomato paste, yogurt, shampoo, beer, wine, milk, and yeast. (ii) Type 2: microbes and microbial lms of bacteria, spores, and yeast species. (iii) Type 3: milk, whey protein concentrate (WPC), cooked SCM, starch, boiled wort, and egg albumin. Some of the research that has considered the inuence of cleaning parameters in owing systems on the removal behavior of deposits is listed in Table 2 to 4. Table 2 details type 1 deposit removal studies, Table 3 details type 2 deposit removal studies, and Table 4 details type 3 deposit removal studies. The cleaned geometry, effect of CIP parameters, and the method of determining cleaning effectiveness are listed in each Table. The effect of ow has been studied both in terms of the Reynolds number (Re) and the surface shear stress. Both may provide further insight into the effect of removal behavior on ow velocity. Milk processing is a large industry and fouling is a signicant problem, as both protein aggregates and minerals are deposited; Burton (1967) classied the proteinaceous deposit seen in pasteurizers as type A and the mineral deposit seen at UHT temperatures as type B. Reviews of dairy fouling research are presented by Changani and others (1997) and Bansal and Chen (2006). Proteins have been identied as a major source of fouling deposits. Fickak and others (2011) found that increasing the protein concentration of whey protein increased the amount of fouling on a pilot-scale heat exchanger. Holding of milk before heating sections has been shown to aggregate -lactoglobulin in the holding sections rather than the heating sections (de Jong and van der Linden 1992). Christian and others (2002) found that increasing the mineral content of whey protein decreased the extent of fouling on a PHE. WPC is often used in research studies to represent a milk fouling deposit, because it is easier to handle and store than milk, and the fouling composition is thus easier to control and replicate. Robbins and others (1999) compared the cleaning of milk and WPC from a PHE. They found that in the pasteurization and UHT sections of the PHE, both materials fouled heavily. However, in the intermediate section, WPC also fouled excessively, whereas milk did

Vol. 12, 2013 r Comprehensive Reviews in Food Science and Food Safety 127

Critical review in fouling and cleaning . . .

Gasket

T Type A fouling on the he gasket v vessel and th Wall fouling Contact t agar d been (s samples had 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 2Some CIP studies of type 1 deposit. Deposit Toothpaste Geometry 1 m long, 2 OD, 316 L ss pipe (horizontal) Effect of ow or w Increase ow velocity (1 to 3 m/s, decrease cleaning time. Effect of temperature Increase temperature (from 20 C), decrease cleaning time (to a point approximately 40 C). (31 to 51 C), removal of shampoo layers faster at higher temperatures as cleaning proceeds. Effect Re Increase Re (4000 to 250000) decreases cleaning time. Cleaning determinant Turbidity reaches 4 ppm. Reference Cole and others (2010)

Shampoo

316 ss plate (350 mm long, 30 mm ID, 18.3 mm ED) (vertical ow cell)

Mustard

glass T-piece (variable depth T, 4 and 6 cm)

(0.14 to 0.47 m/s) higher ow velocity, more efcient removal at the start of cleaning. Increase ow velocity (1 to 1.88 m/s) increase removal rate. Increase w , decrease number of cells. (i) linearly for plastics (ii) as a curve for glass (0.7, 1.5, 2.3 L/min) increase ow rate, the effect of temperature on cleaning time decreases

Visual MSS and spectrophotometry

Pereira and others (2009)

Yeast cells rehydrated (aged 1 h at ambient)

glass, polypropylene, and polystyrene surfaces (210 90 mm long) in horizontal ow cell

Above a certain Reynolds number, the recirculation zone length becomes constant.

Visual

Jensen and others (2007)

Visual

Guillemot and others (2006)

Tomato paste

316 L ss coupons (circular: 26 mm D) horizontal ow cell

(30, 50, and 70 C) increase temperature, decrease the time to remove deposit

(850 to 4800 Re) increase Re, decrease cleaning time

Visual, image analysis, and MHFS

Christian (2004)

OD = outer diameter, ID = inner diameter, ED = equivalent diameter, ss = stainless steel, MSS = mechatronic surface sensor, MHFS = microfoil heat ux sensor.

not. Compositional analysis revealed protein fouling from both materials in the pasteurizer section. Increasing to UHT temperatures revealed milk fouling to become more mineral-based, whereas the WPC fouling remained predominantly protein-based, suggesting comparison of milk fouling and WPC fouling is not wise at UHT temperatures. Yeast can exhibit type 1 (if in contact with glass) and type 2 (if in contact with stainless steel) cleaning behavior. Guillemot and

others (2006) found that yeast cells could be wholly removed from glass using water but that yeast cells had strong adhesion to stainless steel. The wall shear stress required to remove 50% of the attached cells from stainless steel, denoted as w 50% , was 30 Pa, while for plastics w 50% ranged from 1 to 2 Pa. The effect of CIP parameters on the removal of different deposit types is discussed in the following sections. Even though there is clear evidence that different deposit types are removed
C

128 Comprehensive Reviews in Food Science and Food Safety r Vol. 12, 2013

2013 Institute of Food Technologists

Critical review in fouling and cleaning . . .


Table 3Selected CIP studies of type 2 deposit. Deposit Yeast slurry (aged at 30 C, 5 d) Geometry 316 ss coupons (square: 30 30 mm long) in horizontal ow cell Effect of ow or w Increase in ow velocity (0.26 to 0.5 m/s) decrease cleaning time at 50 and 70 C. Limited effect beyond 0.4 m/s at 20 and 30 C. Effect of temperature Increase temperature, decrease cleaning time. Effect of chemical/pH 1% NaOH Cleaning determinant Visual, image analysis, and MHFS Reference Goode and others (2010)

B. cereus spores

316 L ss pipe (20 cm long, 2.37 cm ID) and 2 way valve (entry to exist 18 cm long, 3.5 cm ID) 316 L ss, (210 90 mm long) in horizontal ow cell 304 L ss pipes (15 102 m long, 2.3 102 m ID) (horizontal)

Yeast cells rehydrated (aged 1 h at ambient) B. cereus spores (in milk)

B. cereus spores (in custard)

Progressive-cavity pump (with axial or tangential exit pipe). Tangential was best. In the axial setup, the number of CFU was greater than 10 CFU/cm in the pump body and gaskets.

Increase w , decrease number of cells barely for stainless steel (10%). Increase w (17.45 to 68.95 Pa, that is, 1.61 to 3.29 m/s) decrease number of spores (after 5 min). Contact time was more important in reducing spores.

NaOH 0.5% (w/w) at 60 C, 2200 L/h, up to 30 min. The % residual spores decreased as cleaning time increased.

Agar overlay technique using TTC (spores appear red)

Le Gentil and others (2010)

Visual

Guillemot and others (2006)

Rinsing at 60 C revealed less spores compared to 20 C at the same soaking times.

0.5% w/w of NaOH at 60 C

Agar overlay technique using TTC (spores appear red)

Leli` evre and others (2002)

Prerinse 0.5 m/s (6 min); 0.2% NaOH at 1.5 m/s, 60 C (10 min); intermediate rinse 0.5 m/s (6 min); 0.2% HNO3 at 1.5 m/s, 60 C (10 min); nal rinse 0.5 m/s (6 min).

Agar overlay technique using TTC (spores appear red)

B en ezech and others (2002)

OD = outer diameter, ID = inner diameter, ED = equivalent diameter, ss = stainless steel, MSS = mechatronic surface sensor, MHFS = microfoil heat ux sensor.

from surfaces differently, the approach to cleaning is typically the Product recovery same: At the end of a process, there can be a signicant 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 rst 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 prodAlthough 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 toothCIP 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, nd that prodthe 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:

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 4Selected CIP studies of type 3 deposit. Soil Starch (with phosphorescent tracer molecules) Geometry Effect ow Effect of temperature Effect Re Effect of chemical/pH Cleaning determinant Visual, image analysis Reference Augustin and others (2010)

Continuous and Local cleaning N/A, constant Re > 25000 N/A. Constant abrupt time has a investigated. 0.5%. expansions minimum (ID 26 mm, where the wss expanding shows a from 26 to 38 maximum and mm) vice versa An increase in An increase from Cooked SCM 316 L ss coupons Increase ow An increase in Re (6500 to 0.5 to 1.5% (sweet (square: 30 velocity from temperature 27500) NaOH did not condensed milk) 30 mm long) 0.25 0.5 (40, 60, and revealed a signicantly m/s, decrease 80 C) revealed a decrease in affect cleaning time linear cleaning time cleaning time at all decrease in according to at higher ow temperatures cleaning time Power law. velocities. Increase Re No cleaning at Egg albumin 316 L ss coupons Increase ow 30 C did not (1090 to clean. 0.1 wt% (circular: 26 less 4840) Increase NaOH. mm D) signicant at decreases temperature, Concentration higher cleaning time decrease in 0.25% to 3% chemical con (at 50 C, 0.1 cleaning time. (at 50 C, 2.3 centrations. to 1% NaOH). L/min) However, 50 C removed At 70 C decreases more deposit 0.1%, cleaning time. at 1% NaOH increase Re, Most than 70 C. increase signicant at cleaning time. low ow (0.7 L/min) Increase WPC 316 L ss coupons Limited benet Increasing Re Limited benet temperature (circular: 26 to increase (1090 to to increase (30 to 70 C) mm D) ow velocity 4840) only concentration decrease horizontal at 70 C and benecial at above 0.5%. 1% NaOH. cleaning time ow cell 0.1% NaOH. Benet if at all ow increase ow rates and at low concenchemical contration. centrations (0.7, 1.5, 2.3 L/min, 0.1%, 0.5%, 1% NaOH) WPC 10 cm sections Increasing ow Wall Re 500 to 6500 N/A. Constant of sstubes (6 rate does not temperature investigated. 0.5%. mm ID 0.15 necessarily did not affect As Re 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.

Visual, image analysis and MHFS

Othman and others (2010)

Visual, image analysis, and MHFS

Aziz (2008)

Visual, image analysis, and MHFS

Christian (2004)

Thermal Gillham and resistance using others (1999) MHFS and mass

OD = outer diameter, ID = inner diameter, ED = equivalent diameter, ss = stainless steel, MSS = mechatronic surface sensor, MHFS = Microfoil heat ux sensor, wss = wall shear stress.

0.55 m/s at 50 C. High ow velocity and low temperature in the product recovery stage revealed the fastest cleaning times. The results suggested that the structure of the toothpaste lm after the product recovery stage is important in determining the overall cleaning time. Product recovery can be done by pigging, in which uid is expelled from a system by the pig which could be solid, liquid, or gas. Solid pigs tend to be used in long sections of straight pipe work where complex geometries do not need to be navigated; for example, in crude oil pipelines to remove parafn wax (Guo and others 2005). The use of crushed ice (with a freezing point depressant) in pigging systems has been developed and researched at the Univ. of Bristol to remove starchwater mixes (Quarini 2002). The void fraction of the ice is controlled so that the pig can navigate bends and T-pieces as well as straight pipe work. Application of this technology in the food and beverage industries is currently limited. The ice is expensive to make and store. A company called Aeolus promotes a Whirlwind technology that uses compressed air to remove soft deposits like fruit juice from pipe work with

bends (see www.aeolustech.co.uk). Application of this technology in the food and beverage industry is also limited as the cost of compressed air is considerable. However, use of air in cleaning is likely to increase in the future as water becomes more precious.

The effect of CIP parameters on type 1 removal Schl usser (1976) compared cleaning behavior of 3 type 1 soils; beer, wine, and milk, illustrated in Figure 6. The products themselves were not heated. The cleaning proles of each product were different. Type 1 products can have a complex rheology, but are often shear thinning, that is, they have an effective viscosity that is a function of shear rate. The shear-thinning rheology of yogurt was determined by Henningsson and others (2007) who also studied the use of water to displace the yoghurt. For ow velocities of 0.05 to 0.25 m/s, yogurt was observed and predicted to ow as a plug. If the process was set up so that yogurt ows as a plug, at changeover, the mixing zone between the 2 yogurts would be smaller and yield reduced losses. Prediction of the mixing zone of a HershelBulkley material with and without wall slip at 0.19 m/s

130 Comprehensive Reviews in Food Science and Food Safety r Vol. 12, 2013

2013 Institute of Food Technologists

Critical review in fouling and cleaning . . .


surface. These are gradually eroded away according to zeroorder 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 ow velocity had the biggest impact on shampoo removal from the ow 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 nal cleaning timecreation of a wavy lm in the product removal stage leads to much more rapid removal than if a smooth lm is created. The authors also found that the remaining lm thickness was independent of pipe length, suggesting that removal is uniform throughout the pipe, as also found by Cole and others (2010). Temperature. For cleaning of tomato paste in a ow cell, Christian (2004) found that an increase in temperature decreased the cleaning time by a linear relationship. Both an increase in temperature from 30 to 70 C and in ow velocity from 0.7 to 2.3 L/min decreased cleaning time. Cleaning time decreased by a factor of 6 from the lowest ow rate and temperature to the highest ow rate and temperature. For tomato paste cleaning, it was found that cleaning time was also correlated with Reynolds number (Christian 2004). As the Re was increased from 800 to 4800, the cleaning time (tc ) decreased according to a power law: tc = 2 106 (Re)0.97 . R2 = 0.81. Jackson and Low (1982) found a critical Re of 6300 for cleaning of dried tomato juice from a PHE, below which little deposit was removed. Shampoo was rinsed at 0.14 m/s, at 31 and 51 C, by Pereira and others (2009). After the initial bulk of shampoo was removed from the ow cell, it was found that the removal of shampoo layers occurred faster at higher temperatures. For toothpaste, Cole and others (2010) found that an increase in the water temperature from 20 to 40 C decreased the cleaning time; however, increasing the temperature above 40 C did not decrease cleaning time any further. The same effect may occur when rinsing shampoo; the investigators did not exceed a water temperature of 51 C in their experiments. Cole and others (2010) found that for toothpaste cleaning (from various length scales and diameters), a dimensionless cleaning time, c = tc u /d (where tc is the cleaning time and d is the pipe diameter), could be plotted as a function of Re, as a power law model: c = 9 107 (Re)0.78 ; with a similar t, R2 = 0.84. Palabiyik and others (2012) found that temperature had a greater effect on toothpaste lm removal than ow velocity, and tted the data. Design. The velocity of 1.5 m/s is the ow velocity most often reported to clean pipe lines effectively in industry CIP (EHEDG 1992). This is, however, anecdotal with no theoretical justication (Changani and others 1997; Tamine 2008). In industrial pipe systems, there are, however, more complicated geometries such as bends, valves, and T-pieces. It raises the question: does increasing the ow velocity decrease the cleaning time of other geometries? This gives a better indication of the effect of ow on the cleaning time of a whole system. Jensen and others (2007) lled a variable depth upstand or downstand (also called a T-piece) made from glass with commercially available mustard and rinsed with ambient water. The geometry used in the study is shown in Figure 7 and is in the downstand position. The downstand depth was tested at 4 and 6 cm. The ow velocity was increased from 1 to 1.88 m/s to dene the effect on cleaning the T-piece. Jensen and others (2007) found

Figure 6Cleaning characteristics of 3 type 1 products, beer, red wine, and milk with water (Schl usser 1976).

was also done by Henningsson and others (2007). With wall slip, it was predicted that the material would have a larger plug ow region. However, predicting the ow of a high-viscosity plug or wall layer is very difcult in practice. Flow and wall shear stress. Flow rate has an effect on the removal rate of type 1 materials. The rheology of tomato paste has been represented by the Carreau model (Bayod and others 2008), and the cleaning behavior of tomato paste in a ow cell has been investigated by Christian (2004). At 30 C, it was found that by increasing the cleaning water ow rate from 0.7 to 1.5 to 2.3 L/min (Re 750 and 4840), the cleaning time decreased. The relationship appears linear. This was also true at 50 and 70 C. Shampoo (SUNSILK R color radiant, viscosity quoted as 7000 cP at 24 C) was rinsed by water from a stainless steel plate in a vertical ow cell by Pereira and others (2009), and they found that the faster the initial ow rate (in the range of 0.14 to 0.47 m/s), the more shampoo was removed from a duct. The same effect was found for removing toothpaste (a HershelBulkley uid with a yield stress) from a pipe (Cole and others 2010). The effect of wall shear stress ( w ) in the range of 0.5 to 10 Pa on toothpaste removal was studied. Shear stress is affected by both uid density and Re that are both affected by temperature. Toothpaste cleaning time is governed by 2 removal phases by Cole and others (2010): (i) Core removalwhere most of the product is removed as a slug of product that can be recovered. (ii) Thin-lm removalwhere the remaining annular wall lm of toothpaste is removed. More recently, 3 phases were dened by a further investigation of the effect of product recovery on cleaning of toothpaste using water (Palabiyik and others 2012): (i) Core removalthe rst few seconds (a time comparable to the residence time of the uid in the system) where approximately half the product mass is removed and the remaining toothpaste coats the pipe wall. Also called the product recovery stage. (ii) Film removalfurther product is removed up to about 1000 s according to a process that is 1st order in deposit weight/thickness, leaving a thinner but continuous lm of toothpaste remaining on the pipe wall. (iii) Patch removalgreater than 1000 s, the continuous lm is broken up and only patches of toothpaste are left on the
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 lled with sour milk and/or spores. An area difcult to clean is dened 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 ow 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. B en ezech 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 congurations (i) with an axial exit pipe, where custard was pumped out of the top of the pump body on the same axis as entry, and (ii) with a tangential exit pipe, 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 0.5 m/s for 6 min; 0.2% HNO3 rinse at 1.5 m/s, 60 C, for 10 min; and nal rinse at 0.5 m/s for 6 min. The group found that in the tangential setup, all parts of the pump were cleaned to the same number of CFU/cm, approximately 10 CFU/cm2 . The authors dened a high level of hygiene as counts less than 18 CFU/cm2 . In the axial set up not all components were cleaned to the same level. There was an increased number of CFU/cm2 in the pump body and gaskets (>18 CFU/cm2 ). To clean tanks, spray devices typically called cleaning heads are used. The design of a cleaning head is of paramount importance to be effective in cleaning. There are 2 main choices: (i) Static cleaning headsThese devices spray cleaning uid onto the tank surface from a xed position. The effectiveness of the cleaning head depends on cleaning uid ow rate and the size and pattern of the holes. (ii) Dynamic cleaning headsThese devices spray cleaning uid onto the tank surface using larger pressures, around 5 Bar (resulting in large wall shear stresses and direct impact force), and rotation to ensure full vessel coverage. The effectiveness of the cleaning head depends on the cleaning uid pressure/ow rate to ensure that the preprogrammed pattern is achieved. Examples of commercially available cleaning heads of both types are shown in Figure 9. Increasing the impact force of a jet stream of uid onto a surface can overcome large deposit hydration times and reduce cleaning times. The fraction effect of time, physical action, temperature, and chemical action delivered to the tank by a static cleaning head (spray ball) and a dynamic cleaning head (high-pressure cleaning head) are given in Figure 10 (Tamine 2008). For spray ball cleaning, time is required to achieve deposit removal. Cleaning time is required to achieve product removal using a static head, and mechanical action is required to achieve product removal using a dynamic cleaning head. Dynamic heads enable cleaning behaviors that are less reliant on contact time with the chemical at high temperature. A type 3 soil could be cleaned in a similar time as a type 1 soil. It should, however, be noted that impingement jets from a rotary device or from many small jets in a static device may cause corrosion problems due to rouging, from small iron particles worn from the orices of the thin walled static spray devices that then deposit on the tank wall.
2013 Institute of Food Technologists

Figure 7Downstand geometry used for investigating the inuence of different ow rates during CIP (ow was from left to right) (from Jensen and others 2007).

that: (i) Increasing the ow velocity increased removal rate. However, the authors suggested that this was more likely due to greater acceleration of the water at 1.88 m/s into the T-piece. At the lower ow velocities, ow had not fully developed before entering the T-piece. (ii) Some areas of the T-piece were harder to clean than others. The position in the downstand most difcult to clean was always located in the same position (see Figure 8, shown as a downstand). As expected, the top of the downstand was difcult to clean. However, an additional area located on the downstand pipe was always the last part to be cleaned in all the experiments, regardless of velocity. Jensen and others (2007) used computational uid dynamics (CFD) simulations to predict the wall shear stress in the 4-cm downstand. Their CFD ndings are illustrated in Figure 8(A) to 8(C) where blue is low wall shear stress (0 Pa) and red is high wall shear stress (5 Pa). As the ow velocity was increased, the blue area decreased in size. Within these simulations, the area most difcult to clean, the center of the downstand, is identied. Increasing the ow rate does not improve cleaning of this area. The wall shear stress achieved at this position is low at all 3 ow velocities. The other areas hardest to clean are circled. Jensen and others (2007) examined the effect of pulsed ow in the downstand. They found that pulsing ow only affected the cleaning time of the 4-cm-depth T piece, not the 6-cm-depth T piece. They compared cleaning at 1 m/s (v1 ) and 2 m/s (v2 ) and pulsing at 15 s (p1 ) and 30 s (p2 ). The cleaning time of the 4-cm downstand was longer when rinsed at 1 m/s than when the ow was pulsed. However, rinsing the downstand at 2 m/s gave the quickest cleaning time. The authors concluded that at turbulent Re, the area of the recirculation zone in the T-piece did not change. A recirculation zone is typically located after a pipe expansion and depends on Reynolds number and the expansion ratio. At lower Re (less than 10000 in this case), the length of the recirculation zone may change; hence, cleaning times are shorter for pulsed ow at 1 m/s than using constant ow at 1 m/s. Jensen and Friis (2005) used CFD simulations to predict the cleanability of a mix proof valve fouled with B. stearothermophilus spores in accordance with the EHEDG standard cleanability test (EHEDG 1992; Timperley and others 2000). In the EHEDG

132 Comprehensive Reviews in Food Science and Food Safety r Vol. 12, 2013

Critical review in fouling and cleaning . . .

Figure 8CFD simulations of the ow eld 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).

Figure 9Commercially 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) dened the wetting rate at the mass ow rate (kg/s) required to completely wet a surface of width W (in meter). Wetting rates achieved by single jets from a spray ball were 0.1 to 0.3 kg/ms. The act of removing deposit from a vessel involves initial wetting and subsequent softening (or dissolution) of the deposit, followed by complete removal by further impingement. Morrison and Thorpe (2002) measured the dimensions of the wet area by the impaction of single water jets onto a sheet of painted acrylic for a range of pressures and distances from spray balls of different sized holes. They found that if the jet directly impacted the area to be cleaned, then this area was cleaned within 60 s. The point of impact was smaller than the total area being wetted; however, certain areas were not cleaned by the spray ball. The width of the falling lm from the point of impact remained the same size throughout rinsing. Jet breakup was observed at 45 C, which increased the distribution of the jet and cleaned a larger area. An interesting model for the ow behavior of jets is given by Wilson and others (2012).

The effect of CIP parameters on type 2 and type 3 deposit removal Type 2 deposits can be viscoelastic, temperature-dependent, and/or time-dependent (Rao 1999). Type 3 deposits tend to be thermally induced and precipitate from the process stream onto the heat exchanger surface over time. For example, wort is a complex uid with several components that change structure and solubility upon heating, including carbohydrates, proteins, vitamins, minerC

als, and lipids. The deposits formed during wort boiling are solid and dissimilar to the process stream (Tse and others 2003). Membrane cleaning. There are many types of ltration processes in food and beverage manufacturing operations. The fouling of membranes alters permeability and selectivity and can be characterized by increased pressure differential and decreased membrane ux. Membranes used in the food and bioprocess industries include reverse osmosis (RO), nanoltration (NF), ultraltration (UF), and microltration (MF) (Cui and Muralidhara 2010). In the brewing industry, beer is claried using MF in which yeast readily fouls the membranes. Membrane cleaning is complex as it is necessary to both remove the surface layers and open the pores in the structurethis must be done without the cleaning agent damaging the material. G uell and others (1999) found that when yeast cells were present on cellulose acetate membrane (CAM) as a layer (yeast cake), the yeast was believed to have formed a secondary membrane. Increasing the thickness of the yeast cake reduced permeate ux and protein transmission through the membrane. Increasing the yeast concentration in the feed solution resulted in lower uxes and protein transmission through the CAM. Hughes and Field (2006) discussed the fouling of MF and UF membranes with yeast at subcritical uxes where fouling is negligible. For the MF membrane, the rate of fouling increased with increasing feed concentration, increasing membrane pore size, and decreasing shear stress. The UF membrane could not be cleaned effectively. Mores and Davis (2002) examined the effect of pulsing ow through a CAM to clean it. They found that the ux increased

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 ow 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 ow 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 ow cell, and they found that increasing the ow velocity did not signicantly 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 ow rate having a negligible effect. However, at 70 C, decreased removal efciency was observed, particularly at the highest ow 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 ndings 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 ow cell (Detry and others 2007, 2009) that can, when used correctly, allow ranges of shears to be studied. Le Gentil and others (2010) cleaned Bacillus cereus spores from 316 L stainless steel pipes using 0.5% (w/w) NaOH 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 rst 10 min, up to 70% of the spores were removed, less so in the remaining 20 min. Leli` evre and others (2002) investigated the removal of B. cereus spores from 304 L stainless steel pipes, similar in length and diameter to the pipes used in the study by Le Gentil and others (2010) and 0.5% (w/w) of NaOH was used at 60 C to rinse the pipe. In this study, the effect of ow velocity and temperature was investigated over a 30 min clean. The researchers found that cleaning at 60 C removed more spores than rinsing at 20 C at each 5-min time interval, at the same ow velocity of 1.97 m/s. They found that increasing the ow velocity from 1.61 to 3.29 m/s ( w = 17.45 to 68.95 Pa) at 60 C decreased the number of attached spores in the rst 5 min. However, after this time, the contact time was more important in removing the spores. The increased acceleration at higher ow rates may be controlling the number of spores removed in the rst 5 min of cleaning, as found by Jensen and Friis (2004). Bremer and others (2006) investigated the effect of alkali rinses and acid rinses (formulated and nonformulated) on removing a biolm generated by recirculating skimmed milk powder in a CIP skid for 18 h in 15 mm stainless steel tubes. There were a number
2013 Institute of Food Technologists

Figure 10The fractional importance of different factors: time, coverage, physical action (impact), temperature, and chemical action (chemistry) required for effective tank cleaning by (A) spray ball and (B) rotary jet head (adapted from Tamine 2008).

with increasing shear rate, back pulse pressure, and back pulse duration. At higher shear rate and back pulse pressure, multiple short back pulses were more effective in cleaning the membrane. At low shear rate and back pulse pressure, fewer longer back pulses were more effective. Longer, weaker back pulses led to the highest recovered uxes. Shorrock and Bird (1998) fouled a MF membrane (hydrophilic polyethersulfone, 0.1 m pore diameter), with yeast cake. Water rinsing was found to remove most of the deposit and an increase in temperature from 30 to 60 C was found to decrease fouling resistance (at 0.74 m/s cross-ow velocity (CFV)). At 40 C, using NaOH as optimum concentration, there was optimum ux through the membrane, 0.01% to 0.025%. Formulated sodium hydroxide solution was found to restore membrane ux completely. Cleaning of MF membranes with WPC was considered by Bird and Bartlett (2002) using a at plate stainless steel membrane and by Blanpain-Avet and others (2009) using a tubular ceramic membrane. An optimum alkaline detergent concentration of 0.02% NaOH was found to give maximum ux after cleaning of the stain-

134 Comprehensive Reviews in Food Science and Food Safety r Vol. 12, 2013

Critical review in fouling and cleaning . . .


of conclusions: (i) Rinsing with 1% NaOH (for 10 min, 65 C, 1.5 m/s) followed by 1% nitric acid (for 10 min, 65 C, 1.5 m/s) reduced the number of cells to a similar level than that found after rinsing with only NaOH (at the same conditions). (ii) Formulated detergents (with surfactants, chelating agents, and sequestrants) decreased cell numbers to the same level as rinsing with NaOH (at the same conditions). (iii) Addition of a surface-active agent to the caustic solution signicantly reduced the number of cells compared to standard CIP (NaOH and nitric acid in (i)). (iv) Nitric acid with surfactants removed signicantly more cells than just nitric acid. (v) Addition of a sanitizer step after CIP did not signicantly reduce viable bacteria numbers. This suggests that the concentration of the alkali, the ow velocity, and the temperature could be optimized to give the most efcient cleaning regime where all cells can be removed. Goode and others (2010) investigated the effect of chemical on yeast removal from stainless steel coupons in a ow cell using 2% Advantis 210 (1% NaOH equivalent). They found that increasing the temperature from 20 to 70 C decreased the cleaning time. An increase in ow velocity at 50 and 70 C from 0.26 to 0.5 m/s also decreased the cleaning time; however, at 20 and 30 C, an increase in ow velocity from 0.4 to 0.5 m/s did not signicantly decrease cleaning time. Chemical effects on the cleaning of type 3 deposits. The effect of chemical cleaning of WPC from milk has largely been characterized in the literature as uneven. The cleaning process has 3 distinct phases seen by many independent researchers, for example, by Bird (1992), Gillham (1997), Grasshoff (1997), Tuladhar (2001), and Christian (2004): (i) Swellingalkali solution contacts the deposit and causes swelling, forming a protein matrix of high void fraction. (ii) Erosionuniform removal of deposit by shear stress forces and diffusion. There may be a plateau region of constant cleaning rate, but this depends on the balance between swelling and removal. (iii) Decaythe swollen deposit is thin and no longer uniform so that removal of isolated islands occurs by shear stress and mass transport. Many authors quote 0.5% NaOH to be optimal for WPC removal from stainless steel, although the existence of cleaning optima has not been categorically proved in all cases. Bird and Fryer (1991) found that increasing the NaOH concentration to 2% can produce a deposit with a less open (dissolved) structure than at 0.5%, thus lengthening the swelling phaseYoo and others (2007) and Saikhwan and others (2010) explained the processes that underpin this observation. Plett (1985) reported that a maximum cleaning rate occurs when cleaning with detergent. The contribution of ow rate is hard to determine in chemical cleaning because both shear stress imposed on the deposit and mass ow to the deposit are dependent on the ow rate. In general, the higher the ow rate, the shorter the cleaning time. Timperley and Smeulders (1988) found that the cleaning time of a PHE decreased with increasing ow velocity from 0.2 to 0.5 m/s. There are arguments supporting higher ow rates that create turbulent conditions. This is because turbulent conditions are known to make the ow patterns of the microscopic boundary layer unstable. However, Bird and Fryer (1991) found that there was no signicant change in cleaning rate
2013 Institute of Food Technologists

when moving from laminar to turbulent ow. Disruption of the boundary layer is further discussed in Section 4.1. Generally increasing the temperature decreases the cleaning time. Gillham and others (1999) found that removal of whey protein deposits from stainless steel pipes was strongly dependent on temperature (less so the swelling phase). SCM is an intermediate in the manufacture of some confectionary products, made by evaporating water from milk and adding sugar to lower the water activity of the product. SCM has 70% to 74% total solids of which 40% to 45% is sucrose (Fisher and Rice 1924) leaving 29% to 30% milk solids. In the study of Othman and others (2010), SCM was cooked for 4 h at 85 to 90 C on stainless steel coupons and washed by chemical cleaning in a ow cell. It was found that increasing the ow velocity from 0.25 to 0.5 m/s decreased the cleaning time at all temperatures. An increase in temperature from 40 to 80 C decreased the cleaning time linearly. Interestingly, the authors found that increasing the NaOH concentration from 0.5% to 1.5% did not signicantly affect the cleaning time at each temperature. This agrees with ndings for WPC cleaning that quote 0.5% NaOH as the optimum concentration. It was the increase in temperature rather than the increase in chemical concentration that decreased cleaning time. Cleaning time was plotted compared with Re for SCM at 1% NaOH (40, 60, and 80 C) by Othman and others (2010) and at 0.1%, 0.5%, and 1% NaOH (at 30, 50, and 70 C) for WPC by Christian (2004) as shown in Figure 11(A) and 11(B). For WPC, the range of investigated Re was around 800 to 4840. There were separate groups of data at each temperature that could not be plotted on one master curve. This suggests that temperature was the dominant parameter in controlling cleaning time. Christian (2004) concluded that an increase in Re was only benecial to cleaning time at low concentration. Jennings and others (1957) suggested the existence of a threshold Re of 25000 for cleaning a pipe surface of dry milk deposit before an increase in Re resulted in increased cleaning rate. For SCM, the Re range investigated was much higher, from 6500 to 27000. All the data collapsed onto one curve. As the Re increased, the cleaning time decreased, suggesting that Re was the dominant parameter controlling cleaning time. Othman and others (2010) did nd, however, that the effect of Re on cleaning time became less signicant as the temperature was increased. Gillham and others (1999) found that tc was proportional to Re n , where n was in the range of 0.2 to 0.35 for 0.5% NaOH. For SCM at 1% NaOH, tc was again proportional to Re n where n = 1.28 (R2 = 0.92). The cleaning of egg albumin was characterized by Aziz (2008). Generally, an increase in temperature decreased the cleaning time. However, at 1% NaOH, cleaning time was faster at 50 C than at 70 C. Deposit was not removed at 30 C at any ow velocity or NaOH concentration investigated. An increase in NaOH concentration from 0.25% to 3% NaOH decreased the cleaning time (at 50 C, 2.3 L/min); however, increasing the ow rate had a less signicant impact on cleaning time at higher chemical concentration. The author concluded a high temperature, mid to high ow velocity, and mid-range chemical concentration appeared to be the optimum, similarly to WPC cleaning optima. For egg albumin, the range of Re investigated was 1090 to 4840. There were separate groups of data at 50 and 70 C that could not be plotted on one master curve. This suggests that temperature was the dominant parameter in controlling cleaning time similarly to WPC. In Christian (2004), WPC cleaning experiments were conducted using 0.5% NaOH at 30, 50, and 70 C and 0.7, 1.5, and

Vol. 12, 2013 r Comprehensive Reviews in Food Science and Food Safety 135

Critical review in fouling and cleaning . . .

Figure 12Stationary ow and oscillating components of ow, wos symbolizes an oscillating uid movement, wos,max is maximum oscillating uid velocity, wstat is mean uid velocity, and t is frequency of oscillation (Augustin and others 2010).

an oscillating uid movement, wos , is superimposed, as illustrated in Figure 12. The intensity of a superimposed pulsation (wos,max + wstat ) can be quantied using waviness, W, the ratio of the maximum oscillating (wos,max ) and the stationary or mean ow velocities, w: W= wos,max w (4)

Figure 11Re compared with visual cleaning time of (A) SCM at 40, 60, 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 2004). Shading denotes the ow temperature: Black: 30 C, Gray: 50 C, and Open: 70 C.

w for an oscillation interval is dened by Augustin and others (2010) as: w= 1 tos w (t ) d t with w (t ) = wstat + wos tos 0 = wstat + wos,max . sin(w t )

(5)

2.3 L/min. Rd (the fouling resistance) is a measure of resistance to the ow of heat to the sensor. Rd measured at the same ow rate reduced Rd more rapidly as the temperature was increased from 30 to 70 C. An increase in ow rate from 1.5 to 2.3 L/min revealed similar Rd proles, suggesting that temperature dictated the cleaning time in this case. For all type 3 deposits detailed here, temperature seems to be the dominant contributor to cleaning time at both low and high ow velocity and low and high concentration. Reaction rate, solubility, and possible phase transitions (such as in fats) will all be affected by temperature.

A higher value of waviness is believed to result in a separation of the viscous sublayer and the formation of eddy currents; ow reversal is critical. This effect can decrease the thickness of the laminar sublayer at the pipe surface when applying a turbulent ow. The temporary maximum velocity, wos,max , can occur near the pipe wall resulting in large shear rates and high wall shear stresses. Bode and others (2007) characterized a linear decrease of cleaning time with increasing waviness from 1 to 5. Below a waviness of 1, there is limited ow reversal in the pipe and the cleaning time increases. Augustin and others (2010) compared the cleaning rate of deposit at a ow velocity of 1 m/s (Re 25000), where waviness was 0, and pulsing the ow, where waviness was 1. They found that the amount of deposit in the pipe became asympNovel Cleaning Approaches There have been various methods reported in the literature to totic at 4 min using the pulsed-ow regime and 6 min using the stationary ow regime. Augustin and others (2010) have validated improve cleaning. CFD models accompanying their cleaning data. Phosphorescent Increasing boundary layer disruption zinc sulde crystals were used as an optical tracer enabling accurate Various authors have considered pulsing ow in pipes to en- characterization of ow patterns. hance wall shear stress at lower average ow velocities to enhance cleaning. Gillham and others (2000) showed that pulsing ow at Alternative cleaners to reduce environmental impact a relatively low frequency (2 Hz) enhanced cleaning of a tubuThe efcacy of enzymes in cleaning has been investigated over lar heat exchanger compared to the same steady-ow velocity. A the last decade. Grasshoff (2002) investigated the efcacy of Savpulsed ow creates high periodic accelerations of the liquid ow. inase, a protease (optimum temperature 50 C, pH 9.5), in cleaning The directional change in the ow can increase mass transfer of a milk pasteurizer following a 15-min acid wash. Increasing the the cleaning uid to the surface, thus decreasing cleaning times. concentration of the enzyme from 0.0025% to 0.05% at 60 C A pulsed ow is characterized by a stationary base ow 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
2013 Institute of Food Technologists

Critical review in fouling and cleaning . . .


heat exchanger was opened and inspected after 45 min of enzyme cleaning and 6 min of water rinsing (30 L/min). The interior surfaces were clean. Microbiological product samples collected showed no indication of microbial or enzyme contamination after CIP. The use of commercial enzymes to clean UF membranes has been discussed by Petrus and others (2008) and Allie and others (2003). Petrus and others (2008) used proteases to clean proteins (BSA and -Lg) and dened an optimum concentration of 0.1% for 60 min. The enzyme deposited on the membrane when rinsed for longer than 60 min. Allie and others (2003) used proteases and lipases to clean abattoir efuent. Up to 55% of fouling was removed with lipases, and up to 70% by using lipases and proteases and a ux recovery of up to 100%. However, to apply enzyme cleaning in industry enzyme dosage, process control, and the overall economic burden need to be considered. Orgaz and others (2006) tested the efcacy of fungal enzymes at breaking Pseudomonas uorescens biolm bonds. Out of the tested enzymes, a T. viride enzyme (cultured on pectin) was most effective at removing the biolm by 84% ( 2%). The enzyme was cellulose-pectinesterase-based. The least effective enzyme removed 19% ( 6%) of the biolm and was mostly cellulose. spectively. The micromanipulation technique employs controlled strain parallel to the deposit, where a T-shaped probe is pulled across a horizontal circular plate at a constant height, removing the fouling deposit by a shoveling action. The FDG was developed to measure the thickness of soft-solid meso- and macrolayers on surfaces. The surface is immersed in a Newtonian liquid and a nozzle is brought close to the surface. A suction pressure differential is applied so that liquid is drawn into the nozzle. The gauging suction is increased until deformation is observed. Deformation of the deposit results in changes in h, the distance between the deposit and the gauge tip, which can be readily detected. The 2 techniques showed similar trends and complementary phenomenological detail when removing tomato paste in parallel studies (Hooper and others 2006). As the baking time increased, the adhesive strength increased, and as the hydration time increased, the adhesive strength decreased. The FDG technique has been used to study the effect of alkaline cleaning chemical concentration (0.3% to 2%), temperature (20 to 50 C), and velocity (0.03 to 0.3 m/s, Re: 500 to 10000) on whey protein (Tuladhar and others 2002), the effect of drying time on tomato paste (Chew and others 2004), and the effect of temperature and pH on gelatin lm swelling (Gorden and others 2010). The micromanipulation technique has been used to study the strength of P. uorescens biolm with growth medium velocity (Chen and others 1998), the effect of baking, hydration time, and temperature on tomato paste (Liu and others 2002), and surface energy and cut height on tomato paste (Liu and others 2006). The same study considered the effect of drying/aging and cut height on bread dough, egg albumin, and whey protein. The effect of ovalbumin concentration, temperature, and NaOH concentration on egg albumin adhesion is presented by Liu and others (2007). Tomato paste, bread dough, and egg albumin deposits have been found to have a lower adhesive than cohesive strength, while whey protein was found to have lower cohesive than adhesive strength. Atomic force microscopy (AFM) has also been used to study depositsurface interactions. In this technique, a cantilever contacts the deposit and the force of detaching the cantilever is determined. Beech and others (2002) demonstrated the effective use of AFM for studying the adhesion of bacteria to a wide range of surfaces. Bowen and others (2001) used AFM to study yeast cell detachment from hydrophobic and hydrophilic silicate surfaces with and without protein. The authors found that as yeast cells aged, adhesion force changed. Also, cells in the stationary phase adhered most strongly to the surface. Surface contact was demonstrated to be important where after 5 min, adhesion force increased. Elofsson and others (1997) used AFM to study the coverage of -Lg and WPC to mica sheets at different surface loadings in the range of 0.03 to 3 kg/cm2 . The use of AFM enabled the authors to distinguish between different states of protein aggregation at the submicrometer level.

Other studies related to cleaning behavior Various authors have quantied parameters related to cleaning behavior by other methods rather than in a ow cell or pilot plant. Some techniques used to infer cleaning data are discussed here. Deposit shear. Sim oes and others (2005) fouled stainless steel cylinders with P. uorescens biolm in a bioreactor. Three cylinders were rotated at 500, 1000, 1500, and 2000 min1 sequentially (ReA 2400 to 16100) for 30 s each in phosphate buffer to assess the effect of rotation speed on biolm removal. There was an optimum ReA of 8100 where 45% of biolm was removed. Either side of this ReA removal was less and similar at around 15%. Cylinders were also submerged in different chemical solutions and rotated at 300 min1 for 30 min and then were rotated at 500, 1000, 1500, and 2000 min1 sequentially in phosphate buffer to test cleaning effectiveness. NaOH and sodium hypochlorite (NaClO) representing a CIP detergent and CIP sterilant were investigated. Irrespective of ReA , NaOH and NaClO removed a similar percentage of biolm at the same concentrations (50, 200, and 300 mg/L); however, at the highest concentration, 500 mg/L, NaOH removed approximately 5% more biolm. Demilly and others (2006) characterized the removal of yeast cells as a function of wall shear stress from different stainless steel surfaces using a radial ow chamber, in which ow trajectory was from the center of the plate outwards toward the edges of the plate. This is similar to the impingement of a water jet on a surface. Radial ow investigations may relate to the action of a cleaning head in a tank, to provide a microscale method of investigating water jets, and deposit removal. The steel surfaces were micropolished and electrochemically etched with different sizes of grain and depths. The number of cells remaining from the original number of cells was dened. Interestingly, the authors found a threshold shear stress above which cells were removed regardless of topography; that detachment of yeast cells was faster from etched steel than mirror-polished steel. Deposit deformation and strength. Two methods have been independently developed at Birmingham and Cambridge to determine the strength and deformation behavior of soft-solid fouling layers on hard surfaces immersed in liquid: the micromanipulation technique and the uid dynamic gauging (FDG) technique, reC

Measuring Cleaning
Ensuring good performance of the cleaning process is vital in maintaining reliable product shelf life and quality. Figure 1 demonstrates implementation of a CIP standard in industry, revealing that CIP measurement and control are fundamental to ensuring hygiene on a daily basis. Processes used to establish and run CIP operations include: (i) Validationdetermining the right method of cleaning and setting it as a standard; it is done before implementation of

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 processing plant before cleaning commenced. This effect has been well characterized in pulsed-ow 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 productwater 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 protein gel (HIWPG). The conductivity was seen to increase to 1200 S/cm in the rst 100 to 200 s and decrease thereafter to 200 S/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 measurement 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 rst 200 s and decreased thereafter. At 1200 s, however, the turbidity value increased from 3 to 7 NTU before decreasing further. The turbidity prole contained numerous jagged peaks that may suggest the deposit removal was nonuniform. The prole is quite different from conductivity. Van Asselt and others (2002) tested an inline calcium probe (CHEMFET) during cleaning of a tubular heat exchanger against ofine measurements of calcium (g/L) and pH. During the prerinse, no difference in the online measurement was detected. Ofine measurements revealed a decrease in calcium from 0.5 to 0 g/L in approximately 3 min and an increase in pH from approximately 6 to 8 in 6 min. Conductivity that may have been measured during the prerinse was not presented. Van Asselt and others (2002) also measured turbidity and conductivity online and ofine during cleaning of a dairy evaporator. The evaporator was not prerinsed with water. The conductivity revealed, as expected, an increase during chemical phases and a decrease during water phases. Turbidity determined in-line and ofine gave different deposit removal proles. For example, during the second alkaline rinse online turbidity revealed 2 considerable peaks. This was also seen at the onset of acid to a lesser extent. Ofine measurements did not reveal these peaks. The ofine turbidity samples taken during CIP were analyzed at a later stage which the authors think altered the particle size of the samples, and thus the turbidity measurement. Particles of different sizes scatter light differently, for example, as a function of smaller particle size (Clauberg and Marciniak 2009). However, nonproduct substances (such as air bubbles and detergent) can also absorb light giving misleading data. CellFacts (Coventry, U.K.) has a patented ofine sampling technology which they have demonstrated determines particle size and density in rinse water samples. This was demonstrated in case studies on their website (http://www.cellfacts.com/Cleaning-InPlace.php). The company used this technology to determine particle size and density during CIP of a beer maturation vessel (Scottish & Newcastle Breweries 2008). Malvern Instruments (Malvern, U.K.) have made commercially available technologies that could be used to monitor particle sizes and counts online during CIP of a process line.
2013 Institute of Food Technologists

Figure 13Potential locations of online CIP measurement techniques that could be used in a process line.

a new method and after alterations in an existing operation. It should always be up-to-date. (ii) Vericationare the results correct and accepted? Checks that the system behaves in the predetermined and expected way, after validation. (iii) Monitoringcontinuous monitoring of specic points of a process determining that the process is under control, after verication. When defects in the system have been identied, appropriate action should always be taken as detailed in Figure 1. Typical CIP monitoring tools include visual detection methods, microbial enumeration of CIP rinse water, and in-line probes that measure temperature and the proportion of hydrogen ions (pH) or electrolytes (conductivity) in the cleaning uid. The conductivity of acid and alkali is different to that of water so that the chemical phases of cleaning can be clearly identied and the chemical recovered. The rst step in optimizing CIP is ensuring that the used monitoring techniques are giving reliable data. Yang and others (2008) described the importance of forecasting models in the approach to optimize cleaning. Data obtained at the beginning of CIP (where the condence limit of variation is high) are used to predict the end point of cleaning (where the condence limit of variation is low). Online application of such a tool would be valuable in optimizing CIP. A range of measurement devices have been considered in the literature and have been used to assess either the cleaning uid (bulk measurements) or the surface. Figure 13 illustrates where in a process of line bulk and surface measurements, CIP measurements may be taken. Flow, temperature, conductivity, and pH are bulk measurements.

Online bulk measurements Online bulk measures monitor the uid during cleaning. Typically temperature, ow, pressure, conductivity, and pH are monitored during CIP to ensure that the process is under control. Pressure drop across a system ( P) is dened as: P = PI PO , where PI and PO are the inlet and outlet pressures. PO increases with time during fouling and decreases during cleaning generally. The rate of change of pressure can be used as a measure of cleaning. Fryer and others (2006) illustrated that pressure

138 Comprehensive Reviews in Food Science and Food Safety r Vol. 12, 2013

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. Using the fouling surface as an electrode and a reference electrode on the opposing wall of the test section, the difference in electrical resistance (RE ) across the ow channel can be monitored online (Chen and others 2004). During milk fouling, electrical resistance increased, similar to thermal resistance (Rf ) measurement. During cleaning RE decreased rapidly before any changes in Rf were detected. The authors determined that NaOH chemical penetration 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 implanted in a surface. A measurement series is based on successive voltage pulses of gradually changing amplitude between which the base potential is applied, and the current is continuously measured. Electronic tongues were integrated into a dairy process line and online measurements were taken. A difference in current was observed between the process stream, water, alkali, and acid, although the investigation back to the clean state was not presented. The authors suggested that the dirtiness of the solutions could be distinguished by the tongue.

A 1.5

Exp. 4.14.2 - 50C (Re 2340) Exp. 4.23.1 - 70C (Re 3160)

R d (m K/kW)

1.0

0.5

0.0 -100

400

900

1400

1900

2400

2900

3400

3900

4400

Time (s)

B
2.0

Exp. 5.9- 30C (Re 1600) Exp. 5.33- 50C (Re 2340) Exp. 5.57- 70C (Re 3160)

Rd (m K/kW)

1.5

Online surface measurements Various authors have reported heat transfer measurements during cleaning to assess the effect of cleaning parameters on the heat transfer coefcient and fouling resistance, Rd , including Gillham and others (1999, 2000), Christian (2004), and Aziz (2008). An example of Rd measured during the cleaning of (a) egg albumin and (b) whey protein using 0.5% NaOH at 1.5 L/min is illustrated in Figure 14. Rinsing egg albumin at 30 C did not clean the surface. Deposit remained on the surface even after 3 h of rinsing. Klahre and others (2000) used differential turbidity to monitor biofouling on the pipe walls of water systems. The technique could be considered to study biofouling removal during cleaning. The use of a Mechatronic Surface Sensor (MSS) was being tested to monitor milk components such as calcium phosphate and whey proteins (Pereira and others 2006) and shampoo (Pereira and others 2009). The sensor measures changes in the vibration properties of surfaces due to the buildup or removal of fouling layers. Measuring microbial cleanliness Microbiological enumeration techniques tend to be ofine retrospective techniques. Rinse water samples and surface swabs are plated on selective agar and viable microbes will present themselves within 3 to 7 d, depending on the organism. Most informative swab and plate methods include contact agar method where the agar plate is pressed directly onto the surface and microbes enumerated directly (Salo and others 2008). The cleanliness of a surface can be veried more quickly using ATP bioluminescence that indicates the presence (or absence) of microbes that are alive. However, the measure of ATP does not indicate microbe specicity. Microbes have been identied on surfaces using infrared spectroscopy (Fornalik 2008) and Raman spectroscopy (R osch and others 2003). Visual assessment, image analysis, and mass can all be used to determine the amount of deposit removed (or remaining) on a surface after cleaning; however, these methods are ofine. Janknecht and Melo (2003) published a review discussing online biolm monitoring measurement techniques, known detection limits, and applicability to practical situations. The techC

1.0

0.5

0.0 0 500 1000 1500 2000 2500 3000 3500

Time (s)
Figure 14Rd proles for (A) egg albumin gel (from Aziz 2008) and (B) whey protein (from Christian 2004) with different ow temperatures: 30, 50, and 70 C using 0.5 wt% NaOH and a ow rate of 1.5 L/min.

niques discussed that have been investigated in an industrial setting include:


r Differential turbidimetry in a paper mill (Klahre and others r Light scattering detection (using a ber optic probe) in a brewery r Biolm respiration measurement in a bed reactor (Carri on and

2000),

(Tamachkiarow and Flemming 2003), others 2003).

Goode and others (2010) were able to monitor yeast cleaning from stainless steel coupons using a heat ux sensor and the image analysis technique. Less commercially applied techniques discussed by Janknecht and Melo (2003) include measuring radiation signals (spectroscopy, uorometry, and photoacoustic spectroscopy) and electric and mechanical (vibration) signals. The critical part of optimizing a cleaning process would be the incorporation of cleaning measurements into the process schedule. At present, it is usual that cleaning times are set automatically and are not changed in operation. If measurements made at the start of cleaning could be used to determine the end point of cleaning with high condence, it would be possible to develop some form of a process control strategy that could minimize the cleaning cost (Yang and others 2008). This approach can be successful if measurements are robust, inexpensive, and taken at the dirtiest point within a system.

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


Published information describing the adhesion of fouling materials and microorganisms to a range of surfaces used throughout the food and beverage industries has been presented and discussed. Cleaning operations such as CIP are ubiquitously applied to remove unwanted fouling layers in a processing plant to maintain product safety and process efciency. However, cost benet analysis 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 presentation of recent studies that have investigated the effect of CIP parameters. This has enabled parallels in the literature to be drawn. (i) For type 1 deposits, cleaning time seems to be related to Re. An increase in Re seems to decrease cleaning time according to the power law model. It was also seen that increasing the ow rate or wall shear stress and cleaning temperature to a mid-range temperature (up to 50 C) decreases cleaning time. (ii) For type 2 deposits, water rinsing parameters, temperature, and wall shear stress seemed to have varied effects on removal. Removal behavior seemed to be dependent on the microbial aging time on the surface. NaOH solution removed type 2 deposits in owing systems. When considering one chemical concentration, ow and temperature were seen to have the biggest effect at the start of cleaning, but it was clear that contact time was an important factor as cleaning progressed. (iii) For type 3 deposits, specically protein, an optimum NaOH concentration has been found to occur in numerous studies where excessive chemical material causes formation of a deposit difcult to remove. However, increasing wall shear stress and temperature were most benecial to cleaning. The ndings suggest that optimizing the ow characteristics at a given temperature and concentration is crucial to achieving fast cleaning in all soil cases. This parameter should be optimized in CIP before temperature and or chemical concentration is increased. Novel surface coatings for stainless steel and alternative chemicals for cleaning are being actively researched at an academic scale. However, application of these ndings has not been adopted in industry. The factors affecting the application of research in industry include cost, maintenance, product safety, product quality, and process reliability. The longevity of surface coatings and the traceability of enzymes out of a test system have not been fully demonstrated. However, research in this area is so extensive that it is probably only a matter of time before effective solutions are found. The pulsing of ow to achieve higher wall shear stress within a system looks like a promising route to improve cleanability of process lines. This could be achieved at very low cost because the required equipment is already used in CIP. However, longevity of pumps as a result of pulse cleaning has not been fully determined. In the future, minimizing the water load and environmental impact of cleaning will only become more important. Current issues surrounding novel approaches to cleaning will need to be overcome for application in industrial CIP that will become more important in the future as water becomes less available and/or more expensive. There are a number of methods at various stages of commercialization for monitoring bulk cleaning and cleaning at the surface.

Any probe use to monitor cleaning needs to be robust and of low cost because cleaning cost is believed to be a relatively low cost of the total cost of production. The robustness and applicability of such novel measurement systems again need to be determined in an industrial setting.

References Akhtar N, Bowen J, Asteriadou K, Zhang Z, Fryer PJ. 2010. Matching the nano- to the meso-scale: measuring depositsurface interactions with atomic force microscopy and micromanipulation. Food Bioprod Process 88:3418. Akhtar N, Bowen J, Robbins PT, Fryer PJ, Asteriadou K, Goode KR. 2012. The effect of temperature on adhesion forces between surfaces and model foods containing whey protein and sugar. J Food Eng. submitted. Allie Z, Jacobs EP, Maartens A, Swart P. 2003. Enzymatic cleaning of ultraltration membranes fouled by abattoir efuent. J Membr Sci 218:10716. Asteriadou K, Hasting APM, Bird MR, Melrose J. 2006. Computational uid dynamics for the prediction of temperature proles and hygienic design in the food industry. Food Bioprod Process 84:15763. Augustin W, Fuchs T, F oste H, Sch oler M, Majschak J, Scholl S. 2010. Pulsed ow for enhanced cleaning in food processing. Food Bioprod Process 88:38491. Aziz NS. 2008. Factors that affect cleaning process efciency. [PhD thesis]. School of Chemical Engineering, University of Birmingham. Baier RE. 1980. Substrate inuences on adhesion of microorganisms and their resultant new surface properties. In: Bitton G, Marshall KC, editors. Adsorption of microorganisms to surfaces. New York: John Wiley. p. 59104. Bansal B, Chen XD. 2006. A critical review of milk fouling in heat exchangers. Compr Rev Food Sci Food Safety 5:2733. Bayod E, Willers EP, Tornberg E. 2008. Rheological and structural characterization of tomato paste and its inuence on the quality of ketchup. LWT Food Sci Technol 41:1289300. Beech IB, Smith JR, Steele AA, Penegar I, Campbell SA. 2002. The use of atomic force microscopy for studying interactions of bacterial biolms with surfaces. Colloids Surf B Biointerfaces 23:23147. B en ezech T. 2001. A method for assessing the bacterial retention ability of hydrophobic membrane lters. Trends Food Sci Technol 12:368. B en ezech T, Leli` evre C, Membr e JM, Viet AF, Faille C. 2002. A new test method for in-place cleanability of food processing equipment. J Food Eng 54:715. Bird MR. 1992. Cleaning of food process plant. [PhD thesis]. UK: University of Cambridge. Bird MR, Bartlett M. 2002. Measuring and modelling ux recovery during the chemical cleaning of MF membranes for the processing of whey protein concentrate. J Food Eng 53:14352. Bird MR, Fryer PJ. 1991. An experimental study of the cleaning of surfaces fouled by whey proteins. Food Bioprod Process 69:1321. Blanpain-Avet P, Migdal JF, B en ezech T. 2009. Chemical cleaning of a tubular ceramic microltration membrane fouled with a whey protein concentrate suspensioncharacterization of hydraulic and chemical cleanliness. J Membr Sci 337:15374. Bode K, Hooper RJ, Augustin W, Paterson WR, Wilson DI, Scholl S. 2007. Pulsed ow cleaning of whey protein fouling layers. Heat Transf Eng 28:2029. Bott TR. 1990. Fouling notebook. Rugby, UK: Institution of Chemical Engineers. Bott TR. 1995. Fouling of heat exchangers. New York: Elsevier. Bowen WR, Lovitt R W, Wright CJ. 2001. Atomic force microscopy study of the adhesion of Saccharomyces cerevisiae. J Colloid Interface Sci 237: 5461. Bremer PJ, Fillery S, McQuillan AJ. 2006. Laboratory scale clean-in-place (CIP) studies on the effectiveness of different caustic and acid wash steps on the removal of dairy biolms. Int J Food Microbiol 106:25462. Burton H. 1967. Deposits from whole milk in heat treatment plant: a review and discussion. J Dairy Res 34:13743. Busscher HJ, Norde W, Sharma PK, van der Mei HC. 2010. Interfacial re-arrangement in initial microbial adhesion to surfaces. Curr Opin Colloid Interface Sci 15:5107.

140 Comprehensive Reviews in Food Science and Food Safety r Vol. 12, 2013

2013 Institute of Food Technologists

Critical review in fouling and cleaning . . .


Carri on M, Asaff A, Thalasso F. 2003. Respiration rate measurement in a submerged xed bed reactor. Water Sci Technol 47:2014. Changani SD, Belmar-Beiny MT, Fryer PJ. 1997. Engineering and chemical factors associated with fouling and cleaning in milk processing. Exp Therm Fluid Sci 14:392406. Characklis WG. 1980. Biolm development and destruction. Final Report EPRI CS-1554, project RP902-1, Palo Alto, Calif.: Electric Power Research Institute. Chen MJ, Zhang Z, Bott TR. 1998. Direct measurement of the adhesive strength of biolms in pipes by micromanipulation. Biotechnol Tech 12:87580. Chen M-Y, Chen M-J, Lee P-F, Cheng L-H, Huang l-J, Lai C-H, Huang K-H. 2010. Towards real-time observation of conditioning lm and early biolm formation under laminar ow conditions using a quartz crystal microbalance. Biochem Eng J 53:12130. Chen XD, Li DXY, Lin SXQ, Ozkan N. 2004. On-line fouling/cleaning detection by measuring electric resistanceequipment development and application to milk fouling detection and chemical cleaning monitoring. J Food Eng 61:1819. Chew JYM, Paterson WR, Wilson DI. 2004. Fluid dynamic gauging for measuring the strength of soft deposits. J Food Eng 65:17587. Christian GK. 2004. Cleaning of carbohydrate and dairy protein deposits. [PhD thesis]. U.K.: School of Chemical Engineering, University of Birmingham. Christian GK, Changani SD, Fryer PJ. 2002. The effect of adding minerals on fouling from whey protein concentrate: development of a model fouling uid for a plate heat exchanger. Food Bioprod Process 80:2319. Christian GK, Fryer PJ. 2006. The effect of pulsing cleaning chemicals on the cleaning of whey protein deposits. Food Bioprod Process 84: 3208. Clauberg B, Marcinak M. 2009. Throwing light on photometry. Brewer Distiller Int 5:4850. Cluett JD. 2001. Cleanability of certain stainless steel surface nishes in the brewing process. [Mphil thesis]. South Africa: Faculty of Mechanical Engineering, Rand Afrikaans University. Cole P, Asteriadou K, Robbins PT, Owen EG, Montague GA, Fryer PJ. 2010. Comparison of cleaning of toothpaste from surfaces and pilot scale pipe work. Food Bioprod Process 88:392400. Cui ZF, Muralidhara HS. 2010. Membrane technology: a practical guide to membrane technology and applications in food and bioprocessing. Chapter 10. Oxford, U.K.: Elsevier Science. De Jong P, van der Linden HJLJ. 1992. Design and operation of reactors in the dairy industry. Chem Eng Sci 47:37618. Demilly M, Br echet Y, Bruckert F, Boulang e L. 2006. Kinetics of yeast detachment from controlled stainless steel surfaces. Colloids Surf B Biointerfaces 51:719. Detry JG, Deroanne C, Sindic M, Jensen BBB. 2009. Laminar ow in radial ow cell with small aspect ratios: numerical and experimental study. Chem Eng Sci 64:3142. Detry JG, Rouxhet PG, Boulang e-Petermann L, Deroanne M, Sindic M. 2007. Cleanability assessment of model solid surfaces with a radial-ow cell. Colloids Surf A Physicochem Eng Aspects 302:5408. Detry JG, Sindic M, Deroanne C. 2010. Hygeine and cleanability: a focus on surfaces. Crit Rev Food Sci Nutr 50:583604. Dhadwar SS, Bemman T, Anderson WA, Chen P. 2003. Yeast cell adhesion on oligopeptide modied surfaces. Biotechnol Adv 21:395406. Dror-Ehre A, Adin A, Markovich G, Mamane H. 2010. Control of biolm formation in water using molecularly capped silver nanoparticles. Water Res 44:26019. EHEDG. 1992. A method for assessing the in-place cleanability of food processing equipment. Trends Food Sci Technol 3:3258. EHEDG Yearbook. 2007. Materials of construction for equipment in contact with food Trends Food Sci Technol. 18:S40S50. Elofsson C, Dejmek P, Paulsson M, Burling H. 1997. Atomic force microscopy studies on whey proteins. Int Dairy J 7:8139. Epstein N. 1983. Thinking about heat transfer fouling: a 5 5 matrix. Heat Transf Eng 4:4356. Everaert EPJM, van der Mei HC, Busscher HJ. 1998. Adhesion of yeasts and bacteria to uoro- alkylsiloxane layers chemisorbed on silicone rubber. Colloids Surf B Biointerfaces 10:17990. Fickak A, Al-Raisi A, Chen WD. 2011. Effect of whey protein concentration on the fouling and cleaning of a heat transfer surface. J Food Eng 104:32331.
C

Fisher RC, Rice FE. 1924. Sweetened condensed milk II. A comparative study of methods for determining total solids. J Dairy Sci 7:497502. Fornalik M. 2008. Detecting biofouling in food processing systems. Photonics Food 42:5861. Fryer PJ, Asteriadou K. 2009. A prototype cleaning map: a classication of industrial cleaning processes. Trends Food Sci Technol 20:22562. Fryer PJ, Christian GK, Liu W. 2006. How hygiene happens: physics and chemistry of cleaning. Int J Dairy Technol 59:7684. Gallardo-Moreno AM, Gonz alez-Martna ML, Bruque JM, P erez-Giraldo C. 2004. The adhesion strength of Candida parapsilosis to glass and silicone as a function of hydrophobicity, roughness and cell morphology. Colloids Surf A 249:99103. Garrett TR, Bhakoo M, Zhang Z. 2008. Bacterial adhesion and biolms on surfaces. Prog Nat Sci 18:104956. Geesey GG, Gillis RJ, Avci R, Daly D, Hamilton M, Shope P, Harkin G. 1996. The inuence of surface features on bacterial colonisation and subsequent substratum chemical changes of 316L stainless steel. Corros Sci 38:7395. Gillham CR. 1997. Enhanced cleaning of surfaces fouled by whey protein. [PhD thesis]. U.K.: University of Cambridge. Gillham CR, Fryer PJ, Hasting APM, Wilson DI. 1999. Cleaning-in-place of whey protein fouling deposits: mechanisms controlling cleaning. Food Bioprod Process 77:12738. Gillham CR, Fryer PJ, Hasting APM, Wilson DI. 2000. Enhanced cleaning of whey protein soils using pulsed ows. J Food Eng 46:199209. Goode KR. 2012. Characterising the cleaning behaviour of brewery foulants to minimise the cost of cleaning in place operations. [EngD thesis]. University of Birmingham. Goode KR, Asteriadou K, Fryer PJ, Picksley M, Robbins PT. 2010. Characterising the cleaning mechanisms of yeast and the implications for cleaning in place (CIP). Food Bioprod Process 88:36574. Gordon PW, Brooker ADM, Chew YMJ, Wilson DI, York DW. 2010. Studies into the swelling of gelatine lms using a scanning uid dynamic gauge. Food Bioprod Process 88:35764. Grasshoff A. 1997, Cleaning of heat treatment equipment. IDF Monograph, Fouling and Cleaning in Heat Exchangers, Brussels, Belgium: International Dairy Federation. Grasshoff A. 2002. Enzymatic cleaning of milk pasteurisers. Food Bioprod Process 80:24752. G uell C, Czekaj P, Davis RH. 1999. Microltration of protein mixtures and the effects of yeast on membrane fouling. J Membr Sci 155:11322. Guillemot G, Vaca-Medina G, Martin-Yken H, Vernhet A, Schmitz P, Mercier-Bonin H. 2006. Shear-ow induced detachment of Saccharomyces cerevisiae from stainless steel: inuence of yeast and solid surface properties. Colloids Surf B 49:12635. Guo B, Song S, Chacko J, Ghalambor A. 2005. Offshore pipelines. Chapter 16: pigging operations. Oxford, UK: Elsevier. Henningsson M, Regner M, Ostergren K, Tr ag ardh T, Dejmek P. 2007. CFD simulation and ERT visualization of the displacement of yoghurt by water, on industrial scale. J Food Eng 80:16675. Hilbert LR, Bagge-Ravn D, Kold J, Gram L. 2003. Inuence of surface roughness of stainless steel on microbial adhesion. Int Biodeterior Biodegrad 52:17585. Hooper RJ, Liu W, Fryer PJ, Paterson WR, Wilson DI, Zhang Z. 2006. Comparative studies of uid dynamic gauging and a micromanipulation probe for strength measurements. Food Bioprod Process 84: 3538. Hughes D, Field R W. 2006. Crossow ltration of washed and unwashed yeast suspensions at constant shear under nominally sub-critical conditions. J Membr Sci 280:8998. Jackson AT, Low WM. 1982. Circulation cleaning of a plate heat exchanger fouled by tomato juice: III. The effect of uid ow rate on cleaning efciency. J Food Technol 17:74552. Janknecht P, Melo LF. 2003. Online biolm monitoring. Rev Environ Sci Bio/Technol 2:26983. Jefferson K. 2004. What drives bacteria to form a biolm? FEMS Microbiol Lett 236:16373. Jennings WG, Mckillop AA, Luick JR. 1957. Circulation cleaning. J Dairy Sci 40:14719. Jensen BBB, Friis A. 2005. Predicting the cleanability of mix-proof valves by use of wall shear stress. J Food Process Eng 28:89106. Jensen BBB, Friis A. 2004. Critical wall shear stress for the EHEDG test method. Chem Eng Process 43:83140.

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 changing average velocity. Trends Food Sci Technol 18:S52S63. Klahre J, Flemming M, Flemming H-C. 2000. Monitoring of biofouling in papermill process waters. Water Res 34:365765. Kim JY, Lee JH, Chang IS, Lee JH, Yi CW. 2011. High voltage impulse electric elds: disinfection kinetics and its effect on membrane bio-fouling. Desalination 283:1116. Le Gentil C, Sylla Y, Faille C. 2010. Bacterial re-contamination of surfaces of food processing lines during cleaning in place procedures. J Food Eng 96:3742. Lelieveld HLM, Mostert MS, Holah J. 2005. Handbook of hygiene control in the food industry. EHEDG, 192 208. Cambridge, U.K.: Woodhead. Leli` evre C, Antonini G, Faille C, B en ezech T. 2002. Cleaning-in-place modelling of cleaning kinetics of pipes soiled by bacillus spores assuming a process combining removal and deposition. Food Bioprod Process 80:30511. Liu W, Aziz NA, Zhang Z, Fryer PJ. 2007. Quantication of the cleaning of egg albumin deposits using micromanipulation and direct observation techniques. J Food Eng 78:21724. Liu W, Christian GK, Zhang Z, Fryer PJ. 2002. Development and use of a micromanipulation technique for measuring the force required to disrupt and remove fouling deposits. Food Bioprod Process 80:28691. Liu W, Fryer PJ, Zhang Z, Zhao Q, Liu Y. 2006. Identication of cohesive and adhesive effects in the cleaning of food fouling deposits. Innov Food Sci Emerg Technol 7:2639. Liu M, Li X, Lin R, Nie W, Zhang N, Ling N. 2004. Fouling prevention with uidised particles in evaporation of traditional Chinese medicine extract. China Particuol 2:813. Lorite GS, Rodrigues CM, de Souza AA, Kranz C, Mizaikoff B, Cotta MA. 2011. The role of conditioning lm formation and surface chemical changes on Xylella fastidiosa adhesion and biolm evolution. J Colloid Interface Sci 359:28995. Mercier-Bonin M, Ouazzani K, Schmitz P, Lorthois S. 2004. Study of bioadhesion on a at plate with a yeast/glass model system. J Colloid Interface Sci 271:34250. Mores WD, Davis RH. 2002. Yeast foulant removal by backpulses in crossow microltration. Journal Membr Sci 208:389404. Morison KR, Thorpe RJ. 2002. Liquid distribution from cleaning-in-place sprayballs. Food Bioprod Process 80:2705. Mozes M, Marchal F, Hermesse MP, van Haecht JL, Reuliaux L, Leonard AJ, Rouxhet PG. 1987. Immobilization of microorganisms by adhesion: interplay of electrostatic and nonelectrostatic interactions. Biotechnol Bioeng 30:43950. M uller-Steinhagen H. 2000. Heat exchanger fouling mitigation and cleaning technologies, 128. Rugby, U.K.: Institution of Chemical Engineers. Orgaz B, Kives J, Pedregosa AM, Monistrol IF, Laborda F, SanJos e C. 2006. Bacterial biolm removal using fungal enzymes. Enzyme Microb Technol 40:516. Othman AM, Asteriadou K, Robbins PT, Fryer PJ. 2010. Cleaning of sweet condensed milk deposits on a stainless steel surface. In: Wilson DI, Chew YMJ, editors. Proceedings of the Fouling & Cleaning in Food Processing Conference. Cambridge: Department of Chemical Engineering. p. 17482. Palabiyik I, Olunloyo B, Fryer PJ, Robbins PT. 2012. Flow regimes in the emptying of pipes lled with viscoelastic material. Int J Multiphase Flow. submitted. Parbhu A, Hendy S, Danne M. 2006. Reducing milk protein adhesion rates. A transient surface treatment of stainless steel. Food Bioprod Process 84:2748. Pereira A, Mendes J, Melo LF. 2009. Monitoring cleaning-in-place of shampoo lms using nanovibration technology. Sens Actuators B 136:37682. Pereira A, Rosmaninho R, Mendes J, Melo LF. 2006. Monitoring deposit build-up using a novel mechatronic surface sensor (MSS). Food Bioprod Process 84:36670. Pereni CI, Zhao Q, Liu Y, Abel E. 2006. Surface free energy effect on bacterial retention. Colloids Surf B Biointerfaces 48:1437. Petrus HB, Chen HLV, Norazman N. 2008. Enzymatic cleaning of ultraltration membranes fouled by protein mixture solutions. J Membr Sci 325:78392. Plett EA. 1985. Cleaning of fouled surfaces. In: Lund DB, Plett EA, Sandu C, editors. Fouling and cleaning in food processing. Wisconsin: University of Madison. Quain D, Storg ards E. 2009. The extraordinary world of biolms. Brewer Distiller Int 5:3133. Quarini J. 2002. Ice-pigging to reduce and remove fouling and to achieve clean-in-place. Appl Thermal Eng 22:74753. Rao MA. 1999. Rheology of uid and semisolid foods principles and applications. U.S.A.: Aspen Publishers. Reynolds TB, Fink GR. 2001. Bakers yeast, a model for fungal biolm formation. Science 291:87881. Robbins PT, Elliott BL, Fryer PJ, Belmar MT, Hasting APM. 1999. A comparison of milk and whey fouling in a pilot scale heat exchanger: implications for modelling and mechanistic studies. Food Bioprod Process 77:97106. R osch P, Schmitt M, Keifer W, Popp J. 2003. The identication of microorganisms by micro-Raman spectroscopy. J Mol Struct 661662:3639. Rosmaninho R, Santos O, Nylander T, Paulsson M, Beuf M, Benezech T, Yiantsios S, Andritsos N, Karabelas A, Rizzo G, M uller-Steinhage H, Melo LF. 2007. Modied stainless steel surfaces targeted to reduce fouling evaluation of fouling by milk components. J Food Eng 80: 117687. Saikhwan P, Mercade-Prieto R, Chew YMJ, Gunasekaran S, Paterson WR, Wilson DI. 2010. Swelling and dissolution in cleaning of whey protein gels. Food Bioprod Process 88:37583. Salo S, Friis A, Wirtanen G. 2008. Cleaning validation of fermentation tanks. Food Bioprod Process 86:20410. Santos O, Nylander T, Rosmaninho R, Rizzo G, Yiantsios S, Andritsos N, Karabelas A, M uller-Steinhagen H, Melo L, Boulang e-Petermann L, Gabet ardh C, Paulsson M. 2004. Modied stainless steel surfaces C, Braem A, Tr ag targeted to reduce foulingsurface characterization. J Food Eng 64: 6379. Schl usser, HJ. 1976. Zur Kinetik von Reinigungsvorg angen an festen Ober achen. Brauwissenschaft 29:2638. Scottish & Newcastle Breweries. 2008. CIP Philosophy, Scottish & Newcastle Achiever Database. Sharma A, Garg D, Gupta JP. 1982. Solidication fouling of parafn wax from hydrocarbons. Lett Heat Mass Transf 9:20919. Shorrock CJ, Bird MR. 1998. Membrane cleaning: chemically enhanced removal of deposits formed during cell harvesting. Food Bioprod Process 76:308. Sim oes M, Pereira MO, Vieira MJ. 2005. Effect of mechanical stress on biolms challenged by different chemicals. Water Res 39:514252. Stewart JC, Seiberling DA. 1996. Cleaning in place. Chem Eng 103:7279. Tamine AV. 2008. Cleaning-in-place: dairy, food and beverage operations. Society of Dairy Technology Series. London: Wiley-Blackwell. Tamachkiarow A, Flemming HC. 2003. On-line monitoring of biolm formation in a brewery water pipeline system with a bre optical device. Water Sci Technol 47:1924. Timperley AW, Boution F, B en ezech T, Carpentier B, Curiel GJ, Haugan K, Hofman J. 2000. A method for the assessment of in-place cleanability of food processing equipment. EHEDG. 2nd ed. Chipping, Campden, England: CCFRA Technology Ltd. p. 114. Timperley DA, Smeulders CNM. 1988. Cleaning of dairy HTST plate heat exchangers: optimisation of the single-stage procedure. J Soc Dairy Technol 41:17. Tse KL, Pritchard AM, Fryer PJ. 2003. The rate and extent of fouling in a single-tube wort boiling system. Food Bioprod Process 81:1322. Tuladhar TR. 2001. Development of a novel sensor for cleaning studies. [PhD thesis]. U.K.: University of Cambridge. Tuladhar TR, Paterson WR, Wilson DI. 2002. Investigation of alkaline cleaning-in-place of whey protein deposits using dynamic gauging. Food Bioprod Process 80:199214. Yoo JI, Chen XD, Mercad e-Prieto R, Wilson DI. 2007. Dissolving heat-induced protein gel cubes in alkaline solutions under natural and forced convection conditions. J Food Eng 79:131521. Van Asselt AJ, Van Houwelingen G, Te Giffel MC. 2002. Monitoring system for improving cleaning efciency of cleaning-in-place processes in dairy environments. Food Bioprod Process 80:27680. Vinogradov AV, Winston M, Rupp CJ, Stoodley P. 2004. Rheology of biolms formed from the dental plaque pathogen S. Mutans. Biolms 1:4956. 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

2013 Institute of Food Technologists

Critical review in fouling and cleaning . . .


topography on the ease of bacterial removal. Colloids Surf B Biointerfaces 51:4453. Whitehead KA, Verran J. 2006. The effect of surface topography on the retention of microorganisms. Food Bioprod Process 84:2539. Wilson DI, Le BL, Dao HAD, Lai KY, Morison KR, Davidson JF. 2012. Surface ow and drainage lms created by horizontal impinging liquid jets. Chem Eng Sci 68:44960. Winquist F, Bjorklund R, Krantz-R ulcker C, Lundstr oma C, Ostergren K, Skoglund T. 2005. An electronic tongue in the dairy industry. Sens Actuators B 111112:299304. Xiaokai X, Chongfang M, Yongchang C. 2005. Investigation of the electromagnetic antifouling technology for scale prevention. Chem Eng Technol 28:15405. Yang A, Martin EB, Montague GA, Fryer PJ. 2008. Towards improved cleaning of FMCG plants: a model-based approach. Comput Aided Chem Eng 25:11616. Zhao Q, Liu Y. 2006. Modication of stainless steel surfaces by electroless Ni-P and small amount of PTFE to minimize bacterial adhesion. J Food Eng 72:26672. Zhao Q, Liu Y, Wang C. 2005a. Development and evaluation of electroless Ag-PTFE composite coatings with anti-microbial and anti-corrosion properties. Appl Surf Sci 252:16207. Zhao Q, Liu Y, Wang C, Wang S, M uller-Steinhagen H. 2005b. Effect of surface free energy on the adhesion of biofouling and crystalline fouling. Chem Eng Sci 60:485865. Zhao Q, Wang C, Liu Y, Wang S. 2007. Bacterial adhesion on the metal-polymer composite coatings. Int J Adhes Adhes 27: 8591. Zhao Q, Wang S, M uller-Steinhagen H. 2004. Tailored surface free energy of membrane diffusers to minimize microbial adhesion. Appl Surf Sci 230:3718.

2013 Institute of Food Technologists

Vol. 12, 2013 r Comprehensive Reviews in Food Science and Food Safety 143

Anda mungkin juga menyukai