14.1 Introduction Chapter 13 has outlined the concept of ‘hygienic design’ and ‘hygienic practices’ in controlling the safety of chilled food products. This chapter deals with hygienic practices, specifically those related to cleaning and disinfection. Contamination in food products may arise from four main sources: the constituent raw materials, surfaces, people (and other animals) and the air. Control of the raw materials is addressed elsewhere in this book and is the only non-environmental contamination route. Food may pick up contamination as it is moved across product contact surfaces or if it is touched or comes into contact with people (food handlers) or other animals (pests). The air acts as both a source of contamination, i.e. from outside the processing area, or as a transport medium, e.g. moving contamination from non-product to product contact surfaces. Provided that the process environment and production equipment have been hygienically designed (Chapter 13), cleaning and disinfection (referred to together as ‘sanitation’) are the major day-to-day controls of the environmental routes of food product contamination. When undertaken correctly, sanitation programmes have been shown to be cost-effective and easy to manage, and, if diligently applied, can reduce the risk of microbial or foreign body contamination. Given the intrinsic demand for high standards of hygiene in the production of short shelf-life chilled foods, together with pressure from customers, consumers and legislation for ever-increasing hygiene standards, sanitation demands the same degree of attention as any other key process in the manufacture of safe and wholesome chilled foods. This chapter is concerned with the sanitation of ‘hard’ surfaces only – equipment, floors, walls and utensils – as other surfaces, e.g. protective clothing 14 Cleaning and disinfection J. Holah, Campden and Chorleywood Food Research Association or skin, have been dealt with under personal hygiene (Chapter 13). In this context, surface sanitation is undertaken to: ? remove microorganisms, or material conductive to microbial growth. This reduces the chance of contamination by pathogens and, by reducing spoilage organisms, may extend the shelf-life of some products. ? remove materials that could lead to foreign body contamination or could provide food or shelter for pests. This also improves the appearance and quality of product by removing food materials left on lines that may deteriorate and re-enter subsequent production runs. ? extend the life of, and prevent damage to equipment and services, provide a safe and clean working environment for employees and boost morale and productivity. ? present a favourable image to customers and the public. On audit, the initial perception of an ‘untidy’ or ‘dirty’ processing area, and hence a ‘poorly managed operation’ is subsequently difficult to overcome. 14.2 Sanitation principles Sanitation is undertaken primarily to remove all undesirable material (food residues, microorganisms, foreign bodies and cleaning chemicals) from surfaces in an economical manner, to a level at which any residues remaining are of minimal risk to the quality or safety of the product. Such undesirable material, generally referred to as ‘soil’, can be derived from normal production, spillages, line-jams, equipment maintenance, packaging or general environmental contamination (dust and dirt). To undertake an adequate and economic sanitation programme, it is essential to characterise the nature of the soil to be removed. The product residues are readily observed and may be characterised by their chemical composition, e.g. carbohydrate, fat, protein or starch. It is also important to be aware of processing and/or environmental factors, however, as the same product soil may lead to a variety of cleaning problems dependent primarily on moisture levels and temperature. Generally, the higher the product soil temperature (especially if the soil has been baked) and the greater the time period before the sanitation programme is initiated (i.e. the drier the soil becomes), the more difficult the soil is to remove. Microorganisms can either be incorporated into the soil or can attach to surfaces and form layers or biofilms. There are a number of factors that have been shown to affect attachment and biofilm formation such as the level and type of microorganisms present, surface conditioning layer, substratum nature and roughness, temperature, pH, nutrient availability and time available. Several reviews of biofilm formation in the food industry have been published including Pontefract (1991), Holah and Kearney (1992), Mattila-Sandholm and Wirtanen (1992), Carpentier and Cerf (1993), Zottola and Sasahara (1994), Gibson et al. 398 Chilled foods (1995) and Kumar and Anand, (1998). In general, however, biofilm formation is usually found only on environmental surfaces, and progression of attached cells through microcolonies to extensive biofilm is limited by regular cleaning and disinfection. Gibson et al.(1995) in studies of attached microorganisms in 17 different processing environments, recorded 79% of isolates as Gram negative rods, 8.6% Gram positive cocci, 6.5% Gram positive rods and 1.2% yeast strains. The most common organisms were Pseudomonas, Staphylococcus and Enterobacter spp. Pseudomonads are environmental psychrotrophic organisms that readily attach to surfaces and are common spoilage organsisms in chilled foods. Other common Gram negatives that have been associated with surfaces are coliform organisms that are widely distributed in the environment and may also be indicators of inadequate processing or post process contamination. Staphylococci are associated with human skin and therefore their presence on surfaces may be as a result of transfer from food handlers. In addition, Mettler and Carpentier (1998) studied the microflora associated with the surfaces in milk, meat and pastry sites and concluded that the micro-flora was specific to the processing environment. Bacteria adhering to the food product contact surfaces may be an important source of potential contamination leading to serious hygienic problems and economic losses due to food spoilage. For example, pseudomonads and many other Gram negative organisms detected on surfaces are the spoilage organisms of concern in chilled foods. The survival of organisms in biofilms may be a source of post process contamination, resulting in reduced shelf life of the product. In addition, Listeria monocytogenes has been isolated from a range of food processing surfaces (Walker et al. 1991, Lawrence and Gilmore 1995 and Destro et al. 1996) and is usually looked for in high-risk processing areas via the company environmental sampling plan. Following HACCP principles, if the food processor believes that biofilms are a risk to the safety of the food product, appropriate control steps must be taken. These would include providing an environment in which the formation of the biofilm would be limited, undertaking cleaning and disinfection programmes as required, monitoring and controlling these programmes to ensure their success during their operation and verifying their performance by a suitable (usually microbiological) assessment. Within the sanitation programme, the cleaning phase can be divided up into three stages, following the pioneering work of Jennings (1965) and interpreted by Koopal (1985), with the addition of a fourth stage to cover disinfection. These are described below. 1. The wetting and penetration by the cleaning solution of both the soil and the equipment surface. 2. The reaction of the cleaning solution with both the soil and the surface to facilitate: peptisation of organic materials, dissolution of soluble organics and minerals, emulsification of fats and the dispersion and removal from the surface of solid soil components. Cleaning and disinfection 399 3. The prevention of redeposition of the dispersed soil back onto the cleansed surface. 4. The wetting by the disinfection solution of residual microorganisms to facilitate reaction with cell membranes and/or penetration of the microbial cell to produce a biocidal or biostatic action. Dependent on whether the disinfectant contains a surfactant and the disinfectant practice chosen (i.e. with or without rinsing), this may be followed by dispersion of the microorganisms from the surface. To undertake these four stages, sanitation programmes employ a combination of four major factors as described below. The combinations of these four factors vary for different cleaning systems and, generally, if the use of one energy source is restricted, this short-fall may be compensated for by utilising greater inputs from the others. 1. mechanical or kinetic energy 2. chemical energy 3. temperature or thermal energy 4. time. Mechanical or kinetic energy is used to remove soils physically and may include scraping, manual brushing and automated scrubbing (physical abrasion) and pressure jet washing (fluid abrasion). Of all four factors, physical abrasion is regarded as the most efficient in terms of energy transfer (Offiler 1990), and the efficiency of fluid abrasion and the effect of impact pressure has been described by Anon. (1973) and Holah (1991). Mechanical energy has also been demonstrated to be the most efficient for biofilm removal (Blenkinsopp and Costerton 1991, Wirtanen and Mattila Sandholm 1993, 1994, Mattila-Sandholm and Wirtanen 1992 and Gibson et al. 1999). In cleaning, chemical energy is used to break down soils to render them easier to remove and to suspend them in solution to aid rinsability. At the time of writing, no cleaning chemical has been marketed with the benefit of aiding microorganism removal. In chemical disinfection, chemicals react with microorganisms remaining on surfaces after cleaning to reduce their viability. The chemical effects of cleaning and disinfection increase with temperature in a linear relationship and approximately double for every 10oC rise. For fatty and oily soils, temperatures above their melting point are used, to break down and emulsify these deposits and so aid removal. The influence of detergency in cleaning and disinfection has been described by Dunsmore (1981), Shupe et al. (1982), Mabesa et al. (1982), Anderson et al. (1985) and Middlemiss et al. (1985). For cleaning processes using mechanical, chemical and thermal energies, generally the longer the time period employed, the more efficient the process. When extended time periods can be employed in sanitation programmes, e.g. soak-tank operations, other energy inputs can be reduced (e.g. reduced detergent concentration, lower temperature or less mechanical brushing). 400 Chilled foods Soiling of surfaces is a natural process which reduces the free energy of the system. To implement a sanitation programme, therefore, energy must be added to the soil to reduce both soil particle-soil particle and soil particle-equipment surface interactions. The mechanics and kinetics of these interactions have been discussed by a number of authors (Jennings 1965, Schlussler 1975, Loncin 1977, Corrieu 1981, Koopal 1985, Bergman and Tragardh 1990), and readers are directed to these articles since they fall beyond the scope of this chapter. In practical terms, however, it is worth looking at the principles involved in basic soil removal, as they have an influence on the management of sanitation programmes. Soil removal from surfaces decreases such that the log of the mass of soil per unit area remaining is linear with respect to cleaning time (Fig. 14.1(a)) and thus follows first-order reaction kinetics (Jennings 1965, Schlusser 1975). This approximation, however, is only valid in the central portion of the plot and, in practice, soil removal is initially faster and ultimately slower (dotted line in Fig. 14.1(a)) than that which a first-order reaction predicts. The reasons for this are unclear, though initially, unadhered, gross oil is usually easily removed (Loncin 1977) whilst ultimately, soils held within surface imperfections, or otherwise protected from cleaning effects, would be more difficult to remove (Holah and Thorpe 1990). Routine cleaning operations are never, therefore, 100% efficient, and over a course of multiple soiling/cleaning cycles, soil deposits (potentially including microorganisms) will be retained. As soil accumulates, cleaning efficiency will decrease and, as shown in plot A, Fig. 14.1(b), soil deposits may for a period grow exponentially. The timescale for such soil accumulation will differ for all Fig. 14.1 Soil removal and accumulation. (a) Removal of soil with cleaning time. Solid line is theoretical removal, dotted line is cleaning in practice. (b) Build up of soil (and/or microorganisms); A, without periodic cleans and B, with periodic cleans. (After Dunsmore et al. 1981). Cleaning and disinfection 401 processing applications and can range from hours (e.g. heat exchangers) to typically several days or weeks, and in practice is controlled by the application of a ‘periodic’ clean (Dunsmore et al. 1981). Periodic cleans are employed to return the surface-bound soil accumulation to an acceptable base level (plot B, Fig. 14.1(b)) and are achieved by increasing cleaning time and/or energy input, e.g. higher temperatures, alternative chemicals or manual scrubbing. A typical example of a periodic clean is the ‘week-end clean down’ or ‘bottoming’. 14.3 Sanitation chemicals In many instances, management view the costs of cleaning and disinfection as the price of the chemicals purchased, primarily because this is the only ‘invoice’ that they see. In reality, however, sanitation chemicals are likely to represent approximately only 5% of the true costs, with labour and water costs being the most significant. The purchase of a good quality formulated cleaning product, whilst being initially more expensive, will more than cover its costs by increasing both the standard of clean and cleaning efficiency. Within the sanitation programme it has traditionally been recognised that cleaning is responsible for the removal of not only the soil but also the majority of the microorganisms present. Mrozek (1982) showed a reduction in bacterial numbers on surfaces by up to 3 log orders whilst Schmidt and Cremling (1981) described reductions of 2–6 log orders. The results of work at CCFRA on the assessment of well constructed and competently undertaken sanitation programmes on food processing equipment in eight chilled food factories is shown in Table 14.1. The results suggest that both cleaning and disinfection are equally responsible for reducing the levels of adhered microorganisms. It is important, therefore, not only to purchase quality cleaning chemicals for their soil removal capabilities but also for their potential for microbial removal. Unfortunately no single cleaning agent is able to perform all the functions necessary to facilitate a successful cleaning programme; so a cleaning solution, or detergent, is blended from a range of typical characteristic components: ? water ? surfactants ? inorganic alkalis Table 14.1 Arithmetic and log mean bacterial counts on food processing equipment before and after cleaning and after disinfection Before cleaning After cleaning After disinfection Arithmetic mean 1.32C210 6 8.67C210 4 2.5C210 3 log mean 3.26 2.35 1.14 No. of observations 498 1090 3147 402 Chilled foods ? inorganic and organic acids ? sequestering agents. For the majority of food processing operations it may be necessary, therefore, to employ a number of cleaning products, for specific operations. This requirement must be balanced by the desire to keep the range of cleaning chemicals on site to a minimum so as to reduce the risk of using the wrong product, to simplify the job of the safety officer and to allow chemical purchase to be based more on the economics of bulk quantities. The range of chemicals and their purposes is well documented (Anon. 1991, Elliot 1980, ICMSF 1980, 1988, Hayes 1985, Holah 1991, Koopal 1985, Russell et al. 1982) and only an overview of the principles is given here. Water is the base ingredient of all ‘wet’ cleaning systems and must be of potable quality. Water provides the cheapest readily available transport medium for rinsing and dispersing soils, has dissolving powers to remove ionic-soluble compounds such as salts and sugars, will help emulsify fats at temperatures above their melting point, and, in high-pressure cleaning, can be used as an abrasive agent. On its own, however, water is a poor ‘wetting’ agent and cannot dissolve non-ionic compounds. Organic surfactants (surface-active or wetting agents) are amphipolar and are composed of a long non-polar (hydrophobic or lyophilic) chain or tail and a polar (hydrophilic or lyophobic) head. Surfactants are classified as anionic (including the traditional soaps), cationic, or non-ionic, depending on their ionic charge in solution, with anionics and non-ionics being the most common. Amphipolar molecules aid cleaning by reducing the surface tension of water and by emulsification of fats. If a surfactant is added to a drop of water on a surface, the polar heads disrupt the water’s hydrogen bonding and so reduce the surface tension of the water and allow the drop to collapse and ‘wet’ the surface. Increased wettability leads to enhanced penetration into soils and surface irregularities and hence aids cleaning action. Fats and oils are emulsified as the hydrophilic heads of the surfactant molecules dissolve in the water whilst the hydrophobic end dissolves in the fat. If the fat is surface-bound, the forces acting on the fat/water interface are such that the fat particle will form a sphere (to obtain the lowest surface area for its given volume) causing the fat deposit to ‘roll-up’ and detach itself from the surface. Alkalis are useful cleaning agents as they are cheap, break down proteins through the action of hydroxyl ions, saponify fats and, at higher concentrations, may be bactericidal. Strong alkalis, usually sodium hydroxide (or caustic soda), exhibit a high degree of saponification and protein disruption, though they are corrosive and hazardous to operatives. Correspondingly, weak alkalis are less hazardous but also less effective. Alkaline detergents may be chlorinated to aid the removal of proteinaceous deposits, but chlorine at alkaline pH is not an effective biocide. The main disadvantages of alkalis are their potential to precipitate hard water ions, the formation of scums with soaps, and their poor rinsability. Cleaning and disinfection 403 Acids have little detergency properties, although they are very useful in making soluble carbonate and mineral scales, including hard water salts and proteinaceous deposits. As with alkalis, the stronger the acid the more effective it is; though, in addition, the more corrosive to plant and operatives. Acids are not used as frequently as alkalis in chilled food operations and tend to be used for periodic cleans. Sequestering agents (sequestrants or chelating agents) are employed to prevent mineral ions precipitating by forming soluble complexes with them. Their primary use is in the control of water hardness ions and they are added to surfactants to aid their dispersion capacity and rinsability. Sequestrants are most commonly based on ethylene diamine tetracetic acid (EDTA), which is expensive. Although cheaper alternatives are available, these are usually polyphosphates which are environmentally unfriendly. A general-purpose food detergent may, therefore, contain a strong alkali to saponify fats, weaker alkali ‘builders’ or ‘bulking’ agents, surfactants to improve wetting, dispersion and rinsability and sequestrants to control hard water ions. In addition, the detergent should ideally be safe, non-tainting, non-corrosive, stable, environmentally friendly and cheap. The choice of cleaning agent will depend on the soil to be removed and on its solubility characteristics, and these are summar- ised for a range of chilled products in Table 14.2 (modified from Elliot 1980). Because of the wide range of food soils likely to be encountered and the influence of the food manufacturing site (temperature, humidity, type of equipment, time before cleaning, etc.), there are currently no recognised laboratory methods for assessing the efficacy of cleaning compounds. Food manufacturers have to be satisfied that cleaning chemicals are working appropriately, by conducting suitable field trials. Although the majority of the microbial contamination is removed by the cleaning phase of the sanitation Table 14.2 Solubility characteristics and cleaning procedures recommended for a range of soil types Soil type Solubility characteristics Cleaning procedure recommended Sugars, organic acids, salt Water-soluble Mildly alkaline detergent High protein foods (meat, Water-soluble Chlorinated alkaline detergent poultry, fish) Alkali-soluble Slightly acid-soluble Starchy foods, tomatoes, Partly water-soluble Mildly alkaline detergent fruits Alkali-soluble Fatty foods (fat, butter, Water-insoluble Mildly alkaline detergent; if margarine, oils) Alkaline-soluble ineffective, use strong alkali Heat-precipitated water Water-insoluble Acid cleaner, used on a hardness, milk stone, Alkaline-insoluble periodic basis protein scale Acid-soluble 404 Chilled foods programme, there are likely to be sufficient viable microorganisms remaining on the surface to warrant the application of a disinfectant. The aim of disinfection is therefore to further reduce the surface population of viable microorganisms, via removal or destruction, and/or to prevent surface microbial growth during the inter-production period. Elevated temperature is the best disinfectant as it penetrates into surfaces, is non-corrosive, is non-selective to microbial types, is easily measured and leaves no residue (Jennings 1965). However, for open surfaces, the use of hot water or steam is uneconomic, hazardous or impossible, and reliance is, therefore, placed on chemical biocides. Whilst there are many chemicals with biocidal properties, many common disinfectants are not used in food applications because of safety or taint problems, e.g. phenolics or metal-ion-based products. In addition, other disinfectants are used to a limited extent only in chilled food manufacture and/or for specific purposes, e.g. peracetic acid, biguanides, formaldehyde, glutaraldehyde, organic acids, ozone, chlorine dioxide, bromine and iodine compounds. Of the acceptable chemicals, the most commonly used products are: ? chlorine-releasing components ? quaternary ammonium compounds ? amphoterics ? quaternary ammonium/amphoteric mixtures. Chlorine is the cheapest disinfectant and is available as hypochlorite (or occasionally as chlorine gas) or in slow releasing forms (e.g. chloramines, dichlorodimethylhydantoin). Quaternary ammonium compounds (Quats or QACs) are amphipolar, cationic detergents, derived from substituted ammonium salts with a chlorine or bromine anion and amphoterics are based on the amino acid glycine, often incorporating an imidazole group. In a (CCFRA) survey undertaken of the UK food industry in 1987, of 145 applications of disinfectants 52% were chlorine based, 37% were quaternary ammonium compounds and 8% were amphoterics. Of these biocides there were, respectively, 44, 30 and 8 branded products used. In a (CCFRA) European survey of 1993, the most common disinfectants used in the UK and Scandinavian countries were QACs for open surfaces and peracetic acid and chlorine for closed, liquid handling surfaces. The survey also showed that open surfaces were usually cleaned with alkaline detergents which were foamed and then rinsed with medium pressure water (250psi) and closed systems were CIP cleaned with caustic followed by acidic detergents with a suitable rinse in- between. A survey of the approved disinfectant products in Germany (DVG listed) in 1994 indicated that 36% were QACs, 20% were mixtures of QACs with aldehydes or biguanides and 10% were amphoterics (Knauer-Kraetzl 1994). More recently the synergistic combinations of QACs and amphoterics have been explored in the UK and these compounds are now widely used in chilled food plants. The characteristics of the most commonly used are compared in Table 14.3. The properties of QAC/amphoteric mixes will be similar to their parent compounds with often enhanced microorganism control. Cleaning and disinfection 405 Within the chilled food industry, particularly for mid-shift cleaning and disinfection in high-risk areas, alcohol based products are commonly used. This is primarily to restrict the use of water for cleaning during production as a control measure to prevent the growth and spread of any food pathogens that penetrate the high-risk area barrier controls. Ethyl alcohol (ethanol) and isopropyl alcohol (isopropanol) have bactericidal and virucidal (but not sporicidal) properties (Hugo and Russell 1999), though they are only active in the absence of organic matter i.e. the surfaces need to be wiped clean and then alcohol reapplied. Alcohols are most active in the 60–70% range, and can be formulated into wipe and spray based products. Alcohol products are used on a small, local scale because of their well recognised health and safety issues. The efficacy of disinfectants is generally controlled by five factors: interfering substances (primarily organic matter), pH, temperature, concentra- tion and contact time. To some extent, and particularly for the oxidative biocides, the efficiency of all disinfectants is reduced in the presence of organic matter. Organic material may react chemically with the disinfectant such that it loses its biocidal potency, or spatially such that microorganisms are protected from its effect. Other interfering substances, e.g. cleaning chemicals, may react Table 14.3 Characteristics of some universal disinfectants Property Chlorine QAC Amphoteric Peracetic acid Microorganism control Gram-positive + + + + + + + + Gram-negative + + + + + + + Spores + C0C0 ++ Yeast + + + + + + + + Developed microbial resistance C0 ++ C0 Inactivation by organic matter + + + + + water hardness C0 + C0C0 Detergency properties C0 ++ + C0 Surface activity C0 ++ ++ C0 Foaming potential C0 ++ ++ C0 Problems with taints +/C0C0 C0 +/C0 Stability +/C0C0 C0 +/C0 Corrosion + C0C0 C0 Safety + C0C0 ++ Other chemicals C0 + C0C0 Potential environmental impact + + C0/+ C0/+ C0 Cost C0 ++ ++ + C0 no effect (or problem). + effect. + + large effect. 406 Chilled foods with the disinfectant and destroy its antimicrobial properties, and it is therefore essential to remove all soil and chemical residues prior to disinfection. Disinfectants should be used only within the pH range as specified by the manufacturer. Perhaps the classic example of this is chlorine, which dissociates in water to form HOCl and the OCl ion. From pH 3–7.5, chlorine is predominantly present as HOCl, which is a very powerful biocide, though the potential for corrosion increases with acidity. Above pH 7.5, however, the majority of the chlorine is present as the OCl ion which has about 100 times less biocidal action than HOCl. In general, the higher the temperature the greater the disinfection. For most food manufacturing sites operating at ambient conditions (around 20oC) or higher this is not a problem as most disinfectants are formulated (and tested) to ensure performance at this temperature. This is not, however, the case in the chilled food industry. Taylor et al. (1999) examined the efficacy of 18 disinfectants at both 10oC and 20oC and demonstrated that for some chemicals, particularly quaternary ammonium based products, disinfection was much reduced at 10oC and recommended that in chilled production environments, only products specifically formulated for low-temperature activity should be used. In practice, the relationship between microbial death and disinfectant concentration is not linear but follows a sigmoidal curve. Microbial populations are initially difficult to kill at low concentrations, but as the biocide concentration is increased, a point is reached where the majority of the population is reduced. Beyond this point the microorganisms become more difficult to kill (through resistance or physical protection) and a proportion may survive regardless of the increase in concentration. It is important, therefore, to use the disinfectant at the concentration as recommended by the manufacturer. Concentrations above this recommended level may thus not enhance biocidal effect and will be uneconomic whilst concentrations below this level may significantly reduce biocidal action. Sufficient contact time between the disinfectant and the microorganisms is perhaps the most important factor controlling biocidal efficiency. To be effective, disinfectants must find, bind to and transverse microbial cell envelopes before they reach their target site and begin to undertake the reactions which will subsequently lead to the destruction of the microorganism (Klemperer 1982). Sufficient contact time is therefore critical to give good results, and most general-purpose disinfectants are formulated to require at least five minutes to reduce bacterial populations by five log orders in suspension. This has arisen for two reasons. Firstly five minutes is a reasonable approximation of the time taken for disinfectants to drain off vertical or near vertical food processing surfaces. Secondly, when undertaking disinfectant efficacy tests in the laboratory, a five-minute contact time is chosen to allow ease of test manipulation and hence timing accuracy. For particularly resistant organisms such as spores or moulds, surfaces should be repeatedly dosed to ensure extended contact times of 15–60 minutes. Cleaning and disinfection 407 Ideally, disinfectants should have the widest possible spectrum of activity against microorganisms, including bacteria, fungi, spores and viruses, and this should be demonstrable by means of standard disinfectant efficacy tests. The range of currently available disinfectant test methods was reviewed by Reybrouck (1998) and fall into two main classes, suspension tests and surface tests. Suspension tests are useful for indicating general disinfectant efficacy and for assessing environmental parameters such as temperature, contact time and interfering matter such as food residues. In reality however, microorganisms disinfected on food contact surfaces are those that remain after cleaning and are therefore likely to be adhered to the surface. A surface test is thus more appropriate. A number of authors have shown that bacteria attached to various surfaces are generally more resistant to biocides than are organisms in suspension (Dhaliwal et al. 1992, Frank and Koffi 1990, Holah et al. 1990a, Hugo et al. 1985, Le Chevalier et al. 1988, Lee and Frank 1991, Ridgeway and Olsen 1982, Wright et al. 1991, Andrade et al. 1998, Das et al. 1998). In addition, cells growing as a biofilm have been shown to be more resistant (Frank and Koffi 1990, Lee and Frank 1991, Ronner and Wong 1993). The mechanism of resistance in attached and biofilm cells is unclear but may be due to physiological differences such as growth rate, membrane orientation changes due to attachment and the formation of extracellular material which surrounds the cell. Equally, physical properties may have an effect e.g. protection of the cells by food debris or the material surface structure or problems in biocide diffusion to the cell/material surface. To counteract such claims of enhanced surface adhered resistance, it can be argued that in reality, surface tests do not consider the environmental stresses the organisms may encounter in the processing environment prior to disinfection (action of detergents, variations in temperature and pH and mechanical stresses) which may affect susceptibility. Both suspension and surface tests have limitations, however, and research based methods are being developed to investigate the effect of disinfectants against adhered microorganisms and biofilms in-situ and in real time. Such methods have been reviewed by Holah et al. (1998). In Europe, CEN TC 216 is currently working to harmonise disinfectant testing and has produced a number of standards. The current food industry disinfectant test methods of choice for bactericidal and fungicidal action in suspension are EN 1276 (Anon. 1997) and EN 1650 (Anon. 1998a) respectively and food manufacturers should ensure that the disinfectants they use conform to these standards as appropriate. A harmonised surface test is expected in 2000. Because of the limitations of disinfectant efficacy tests, however, food manufacturers should always confirm the efficacy of their cleaning and disinfection programmes by field tests either from evidence supplied by the chemical company or from in-house trials. As well as having demonstrable biocidal properties, disinfectants must also be safe (non-toxic) and should not taint food products. Disinfectants can enter food products accidentally e.g. from aerial transfer or poor rinsing, or deliberately e.g. from ‘no rinse status’ disinfectants. The practice of rinsing or 408 Chilled foods not rinsing has yet to be established. The main reason for leaving disinfectants on surfaces is to provide an alleged biocide challenge (this has not been proven) to any subsequent microbial contamination of the surface. It has been argued, however, that the low biocide concentrations remaining on the surface, especially if the biocide is a QAC, may lead to the formation of resistant surface populations. In Europe, legislation is confusing surrounding whether or not disinfectants can be left on surfaces without rinsing. The Meat Products Directive (95/68/EC) allows disinfectants to remain on surfaces (no rinse status) ‘when the directions for use of such substances render such rinsing unnecessary’, whilst the Egg Products (89/437/ EEC) and Milk Products (92/46/EEC) Directives require that disinfectants must be rinsed off by potable water. There is no specific guidance for other food product categories although the general Directive on the hygiene of foodstuffs (93/43/EEC) requires ‘Food business operators shall identify any step in their activities which is critical to ensuring food safety and ensure that adequate safety procedures are identified, implemented, maintained and reviewed. . .’. In terms of the demonstration of non-toxicity, legislation will vary in each country although in Europe, this will be clarified with the implementation of Directive 98/9/EC concerning the placing of biocidal products on the market, which contains requirements for toxicological and metabolic studies. A recognised acceptable industry guideline for disinfectants is a minimum acute oral toxicity (with rats) of 2,000 mg/Kg bodyweight. Approximately 30% of food taint complaints are thought to be associated with cleaning and disinfectant chemicals and are described by sensory scientists as ‘soapy’, ‘antiseptic’ or ‘disinfectant’ (Holah 1995). CCFRA have developed two taint tests in which foodstuffs which have and have not been exposed to disinfectant residues are compared by a trained taste panel using the standard triangular taste test (Anon. 1983a). For assessment of aerial transfer, a modification of a packaging materials odour transfer test is used (Anon. 1964) in which food products, usually of four types (high moisture e.g. melon, low moisture e.g. biscuit, high fat e.g. cream, high protein e.g. chicken) are held above disinfectant solution of distilled water for 24 hours. To assess surface transfer, a modification or a food container transfer test is used (Anon. 1983b) in which food products are sandwiched between two sheets of stainless steel and left for 24 hours. Disinfectants can be sprayed onto the stainless steel sheets and drained off, to simulate no rinse status, or can be rinsed off prior to food contact. Control sheets are rinsed in distilled water only. The results of the triangular test involve both a statistical assessment of any flavour differences between the control and disinfectant treated sample and a description of any flavour changes. 14.4 Sanitation methodology Cleaning and disinfection can be undertaken by hand using simple tools, e.g. brushes or cloths (manual cleaning), though as the area of open surface requiring Cleaning and disinfection 409 cleaning and disinfection increases, specialist equipment becomes necessary to dispense chemicals and/or provide mechanical energy. Chemicals may be applied as low pressure mists, foams or gels whilst mechanical energy is provided by high and low pressure water jets or water or electrically powered scrubbing brushes. These techniques have been well documented (Anon. 1991, Marriott 1985, Holah 1991) and this section considers their use in practice. The use of cleaning techniques can perhaps be described schematically following the information detailed in Fig. 14.2. The figure details the different energy source inputs for a number of cleaning techniques and shows their ability to cope with both low and high (dotted line) levels of soiling. For the manual cleaning of small items a high degree of mechanical energy can be applied directly where it is needed and with the use of soak tanks (or clean-out-of-place techniques) contact times can be extended and/or chemical and temperature inputs increased such that all soil types can be tackled. Alternatively, dismantled equipment and production utensils may undergo manual gross soil removal and then be cleaned and disinfected automatically in tray or tunnel washers. As with soak tank operations, high levels of chemical and thermal energy can be used to cope with the majority of soils. The siting of tray washes in high-risk chilled production areas should be carefully considered, however, as they are prone to microbial aerosol production which may lead to aerial product contamination (see Chapter 13). In manual cleaning of larger areas, for reasons of operator safety, only low levels of temperature and chemical energy can be applied, and as the surface area requiring cleaning increases, the technique becomes uneconomic with respect to time and labour. Labour costs amount to 75% of the total sanitation programme and for most food companies, the cost of extra staff is prohibitive. Only light levels of soiling can be economically undertaken by this method. The main difference between the mist, foam and gel techniques is in their ability to maintain a detergent/soil/surface contact time. For all three techniques, mechanical energy can be varied by the use of high or low-pressure water rinses, though for open surface cleaning, temperature effects are minimal. Mist spraying is undertaken using small hand-pumped containers, ‘knapsack’ sprayers or pressure washing systems at low pressure. Misting will only ‘wet’ vertical smooth surfaces; therefore only small quantities can be applied and these will quickly run off to give a contact time of five minutes or less. Because of the nature of the technique to form aerosols that could be an inhalation hazard, only weak chemicals can be applied, and so misting is useful only for light soiling. On cleaned surfaces, however, misting is the most commonly used method for applying disinfectants. Foams can be generated and applied by the entrapment of air in high-pressure equipment or by the addition of compressed air in low-pressure systems. Foams work on the basis of forming a layer of bubbles above the surface to be cleaned which then collapses and bathes the surface with fresh detergent contained in the bubble film. The critical element in foam generation is for the bubbles to collapse at the correct rate: too fast and the contact time will be minimal; too 410 Chilled foods Fig. 14.2 Relative energy source inputs for a range of cleaning techniques. (Modified from Offiler 1990). slow and the surface will not be wetted with fresh detergent. Gels are thixotropic chemicals which are fluid at high and low concentrations but become thick and gelatinous at concentrations of approximately 5–10%. Gels are easily applied through high- and low-pressure systems or from specific portable electric pumped units and physically adhere to the surface. Foams and gels are more viscous than mists, are not as prone to aerosol formation and thus allow the use of more concentrated detergents, and can remain on vertical surfaces for much longer periods (foams 10–15 minutes, gels 15 minutes to an hour or more). Foams and gels are able to cope with higher levels of soils than misting, although in some cases rinsing of surfaces may require large volumes of water, especially with foams. Foams and gels are well liked by operatives and management, because of the nature of the foam, a more consistent application of chemicals is possible and it is easier to identify areas that have been ‘missed’. Fogging systems have been traditionally used in the chilled food industry to create and disperse a disinfectant aerosol to reduce airborne microorganisms and to apply disinfectant to difficult to reach overhead surfaces. The efficacy of fogging was recently examined in the UK and has been reported (Anon. 1998b). Providing a suitable disinfectant is used, fogging is effective at reducing airborne microbial populations by 2–3 log orders in 30–60 minutes. Fogging is most effective using compressed air driven fogging nozzles producing particles in the 10–20 micron range. For surface disinfection, fogging is effective only if sufficient chemical can be deposited onto the surface. This is illustrated in Figure 14.3 which shows the log reductions achieved on horizontal, vertical and upturned (underneath) surfaces arranged at five different heights from just below the ceiling (276 cm) to just above the floor (10 cm) within a test room. It can be seen that disinfection is greatest on surfaces closest to the floor and that disinfection is minimal on upturned surfaces close to the ceiling. To reduce inhalation risks, sufficient time (45–60 minutes) is required after fogging to Fig. 14.3 Comparative log reductions of microorganisms adhered to surfaces and positioned at various heights and orientations. 412 Chilled foods allow the settling of disinfectant aerosols before operatives can re-enter the production area. Cleaning chemicals are removed from surfaces by low-pressure/high-volume hoses operating at mains water pressure or by high-pressure/low volume pressure washing systems. Pressure washing systems typically operate at between 25–100 bar through a 15o nozzle and may be mobile units, wall mounted units or centralised ring-mains. Water jets confer high mechanical energy, can be used on a wide range of equipment and environmental surfaces, will penetrate into surface irregularities and are able to mix and apply chemicals. Mechanical scrubbers include traditional floor scrubbers, scrubber/driers (automats) for floors, and water-driven attachments to high pressure systems and electrically operated small-diameter brushes that can be used on floors, walls and other surfaces. Contact time is usually limited with these techniques (though can be increased), but the combination of detergency with high mechanical input allows them to tackle most soil types. The main limitation is that food- processing areas have not traditionally been designed for their use, though this can be amended in new or refurbished areas. The hygienic implications of the design and use of cleaning equipment should be carefully considered. Sanitation equipment should be constructed out of smooth, non-porous, easily cleanable materials such as stainless steel or plastic. Mild steel or other materials subject to corrosion may be used but must be suitably painted or coated, whilst the use of wood is unacceptable. Frameworks should be constructed of tubular or box section material, closed at either end and properly jointed, e.g. welds should be ground and polished and there should be no metal-to-metal joints. Crevices and ledges where soil could collect should be avoided and exposed threads should be covered or dome nuts used. Tanks for holding cleaning chemicals or recovered liquids should be self- draining, have rounded corners and should be easily cleaned. Shrouds around brush heads or hoods and rotary scrubbing heads should be easily detachable to facilitate cleaning. Brushes should have bristles of coloured, impervious material, e.g. nylon, embedded into the head with resin so no soil trap points are apparent. Alternatively, brushes with the head and bristles moulded as one unit may be used. Cleaning equipment is prone to contamination with Listeria spp. and other pathogenic microoganisms and, by the nature of its use, provides an excellent way in which contamination can be transferred from area to area. Cleaning equipment should be specific to high risk and after use, equipment should be thoroughly cleaned and, if appropriate, disinfected and dried. The potential for cleaning equipment to disperse microbial contamination by the formation of aerosols has been reported (Holah et al. 1990b) and it was shown that all cleaning systems tested produced viable bacterial aerosols from test surfaces contaminated with attached biofilms. The degree of contamination impinging on a surface was graded from total coverage to the minimum level thought likely to give concern if a proportion of the droplets contained viable microorganisms and the maximum height and Cleaning and disinfection 413 distance travelled by this contamination level is shown in Table 14.4. Assuming an average food contact surface height of 1 m, the results suggest that both the high pressure low-volume (HPLV) and low pressure high-volume (LPHV) techniques disperse a significant level of aerosol to this height and should not, therefore, be used during production periods. The other techniques, however, are acceptable for use in clean-as-you-go operations as the chance of contamination to product is low, though care is needed when using floor scrubber/driers (these are useful in that the cleaning fluid is removed from the floor) if product is stored in racks close to the floor. After production, HPLV and LPHV techniques may be safely used (and are likely to be the appropriate choice), but it is required that disinfection of food contact surfaces is the last operation to be performed within the sanitation programme. Subsequent work has shown that reducing water pressure or changing impact angle made little difference to the degree of aerosol spread for HPLV and LPHV systems, dispersal to heights C621 m still being achieved. 14.5 Sanitation procedures Sanitation procedures are concerned with both the stage at which the sanitation programme is implemented and the sequence in which equipment and environmental surfaces are cleaned and disinfected within the processing area. Sanitation programmes are so constructed as to be efficient with water and chemicals, to allow selected chemicals to be used under their optimum conditions, to be safe in operation, to be easily managed and to reduce manual labour. In this way an adequate level of sanitation will be achieved, economically and with due regard to environmental friendliness. The principal stages involved in a typical sanitation programme are described below. Production periods. Production staff should be encouraged to consider the implications of production practices on the success of subsequent sanitation programmes. Product should be removed from lines during break periods and this may be followed by manual cleaning, usually undertaken by wiping with Table 14.4 Maximum height and distance of aerosol impingement for a number of cleaning techniques Cleaning technique Height (cm) Distance (cm) High-pressure/low-volume spray lance 309 700 Low-pressure/high-volume hose 210 350 Floor scrubber/drier 47 80 Manual brushing 24 75 Manual wiping 23 45 414 Chilled foods alcohol (to avoid the use of water during production periods). Production staff should also be encouraged to operate good housekeeping practices (this is also an aid to ensuring acceptable product quality) and to leave their work stations in a reasonable condition. Soil left in hoppers and on process lines, etc. is wasted product! Sound sanitation practices should be used to clean up large product spillages during production. Preparation. As soon as possible after production, equipment should be dismantled as far as is practicable or necessary to make all surfaces that microorganisms could have adhered to during production accessible to the cleaning fluids. All unwanted utensils/packaging/equipment should be covered or removed from the area. Dismantled equipment should be stored on racks or tables, not on the floor! Machinery should be switched off, at the machine and at the power source, and electrical and other sensitive systems protected from water/chemical ingress. Preferably, production should not occur in the area being cleaned, but in exceptional circumstances if this is not possible, other lines or areas should be screened off to prevent transfer of debris by the sanitation process. Gross soil removal. Where appropriate, all loosely adhered or gross soil should be removed by brushing, scraping, shovelling or vacuum, etc. Wherever possible, soil on floors and walls should be picked up and placed in suitable waste containers rather than washed to drains using hoses. Pre-rinse. Surfaces should be rinsed with low pressure cold water to remove loosely adhered small debris. Hot water can be used for fatty soils, but too high a temperature may coagulate proteins. Cleaning. A selection of cleaning chemicals, temperature and mechanical energy is applied to remove adhered soils. Inter-rinse. Both soil detached by cleaning operations and cleaning chemical residues should be removed from surfaces by rinsing with low pressure cold water. Disinfection. Chemical disinfectants (or occasionally heat) are applied to remove and/or reduce the viability of remaining microorganisms to a level deemed to be of no significant risk. In exceptional circumstances and only when light soiling is to be removed, it may be appropriate to combine stages 5–7 by using a chemical with both cleaning and antimicrobial properties (detergent-sanitiser). Post-rinse. Disinfectant residues should be removed by rinsing away with low pressure cold water of known potable quality. Some disinfectants, however, are intended to be left on surfaces until the start of subsequent production periods and are thus so formulated to be both surface-active and of low risk, in terms of taint or toxicity, to foodstuffs. Inter-production cycle conditions. A number of procedures may be undertaken, including the removal of excess water and/or equipment drying, to prevent the Cleaning and disinfection 415 growth of microorganisms on production contact surfaces in the period up until the next production process. Alternatively, the processing area may be evacuated and fogged with a suitable disinfectant. Periodic practices. Periodic practices increase the degree of cleaning for specific equipment or areas to return them to acceptable cleanliness levels. They include weekly acidic cleans, weekend dismantling of equipment, cleaning and disinfection of chillers and sanitation of surfaces, fixtures and fittings above two metres. A sanitation sequence should be established in a processing area to ensure that the applied sanitation programme is capable of meeting its objectives and that cleaning programmes, both periodic and for areas not cleaned daily, are implemented on a routine basis. In particular, a sanitation sequence determines the order in which the product contact surfaces of equipment and environmental surfaces (walls, floors, drains etc.) are sanitised, such that once product contact surfaces are disinfected, they should not be re-contaminated. Based on industrial case studies, the following sanitation sequence for chilled food production areas, has been demonstrated to be useful in controlling the proliferation of undesirable microorganisms. The sequence must be performed at a ’room’ level such that all environmental surfaces and equipment in the area are cleaned at the same time. It is not acceptable to clean and disinfect one line and then move onto the next and start the sequence again as this merely spreads contamination around the room. 1. Remove gross soil from production equipment. 2. Remove gross soil from environmental surfaces. 3. Rinse down environmental surfaces (usually to a minimum of 2 m in height for walls). 4. Rinse down equipment and flush to drain. 5. Clean environment surfaces, usually in the order of drains, walls then floors. 6. Rinse environmental surfaces. 7. Clean equipment. 8. Rinse equipment. 9. Disinfect equipment and rinse if required. 10. Fog (if required) 14.6 Evaluation of effectiveness of sanitation systems Assessment of the effectiveness of the sanitation programme’s performance is part of day-to-day hygiene testing and, as such, is linked to the factory environmental sampling plan. The control of the environmental routes of contamination is addressed via the development of a thorough risk analysis and management strategy, typically undertaken as part of the factory HACCP study, resulting in the development of the factory environmental sampling plan. The 416 Chilled foods development of environmental sampling plans has recently been established by a CCFRA industrial working party and is reported in Holah (1998). Environmental sampling is directly linked with both process development and product manufacture and as such, has three distinct phases; 1. process development to determine whether a contamination route is a risk and assessing whether procedures put in place to control the risk identified are working 2. routine hygiene assessment 3. troubleshooting to identify why products (or occasionally environmental samples) may have a microbiological count that is out of specification or may contain pathogens. Related to chilled food manufacture, routine hygiene testing is concerned with assessing the performance of the high-risk barrier systems in preventing pathogen access during production and, after production has finished, the performance of the sanitation programme. Routine hygiene testing is an important aspect of due diligence and is used for two purposes, monitoring to check sanitation process control, and verification to assess sanitation programme success. Monitoring is a planned sequence of observations or measurements to ensure that the control measures within the sanitation programme are operating within specification and are undertaken in a time frame that allows sanitation programme control. Verification is the application of methods in a longer time frame to determine compliance with the sanitation programme’s specification. Monitoring the sanitation programme is via physical, sensory and rapid chemical hygiene testing methods. Microbiological testing procedures are never fast enough to be used for process monitoring. Physical tests are centred on the critical control measures of the performance of sanitation programmes and include, for example, measurement of detergent/disinfectant contact time; rinse water, detergent and disinfectant temperatures; chemical concentrations; surface coverage of applied chemicals; degree of mechanical or kinetic input; cleaning equipment maintenance and chemical stock rotation. Sensory evaluation is usually undertaken after each of the sanitation programme stages and involves visual inspection of surfaces under good lighting, smelling for product or offensive odours, and feeling for greasy or encrusted surfaces. For some product soils, residues can be more clearly observed by wiping the surface with paper tissues. Rapid hygiene methods are defined as monitoring methods whose results are generated in a time frame (usually regarded as within approximately 10 minutes) sufficiently quickly to allow process control. Current methodology allows the quantification of microorganisms (ATP), food soils (ATP, protein) or both (ATP). No technique is presently available which will allow the detection of specific microbial types within this time frame. The most popular and established rapid hygiene monitoring technique is that based on the detection of adenosine triphosphate (ATP) by bioluminescence and Cleaning and disinfection 417 is usually referred to as ATP testing. ATP is present in all living organisms, including microorganisms (microbial ATP), in a variety of foodstuffs and may also be present as free ATP (usually referred to together as non-microbial ATP). The bioluminescent detection system is based on the chemistry of the light reaction emitted from the abdomen of the North American firefly Photinus pyralis, in which light is produced by the reaction of luciferin and luciferase in the presence of ATP. For each molecule of ATP present, one photon of light is emitted which are then detected by a luminometer and recorded as relative light units (RLU). The reaction is very rapid and results are available within seconds of placing the sample to be quantified in the luminometer. The result, the amount of light produced, is also directly related to the level of microbial and non-microbial ATP present in the sample and is often referred to as the ‘hygienic’ status of the sample. ATP has been successfully used to monitor the hygiene of surfaces for approximately 15 years and many references are available in the literature citing its proficiency and discussing its future potential e.g. Bautista et al. (1992), Poulis et al. (1993), Bell et al.(1994), Griffiths et al. (1994), Hawronskyj and Holah (1997). It is possible to differentiate between the measurement of microbial and non-microbial ATP but for the vast majority of cases, the measurement of total ATP (microbial and non-microbial) is preferred. As there is more inherent ATP in foodstuffs than in microorganisms, the measurement of total ATP is a more sensitive technique to determine remaining residues. Large quantities of ATP present on a surface after cleaning and disinfection, regardless of their source, is an indication of poor cleaning and thus contamination risk (from microorganisms or materials that may support their growth). Many food processors typically use the rapidity of ATP to allow monitoring of the cleaning operation such that if a surface is not cleaned to a predetermined level it can be recleaned prior to production. Similarly, pieces of kit can be certified as being cleaned prior to use in processing environments where kit is quickly recycled or when the manufacturing process has long production runs. Some processors prefer to assess the hygiene level after the completion of both the cleaning and disinfection phases whilst others monitor after the cleaning phase and only go onto the disinfection phase if the surfaces have been adequately cleaned. Techniques have also been developed which use protein concentrations as markers of surface contamination remaining after cleaning operations. As these are dependent on chemical reactions, they are also rapid but their applicability is perhaps less widespread as they can only be used if protein is a major part of the food product processed. As with the ATP technique, a direct correlation between the degree of protein remaining after a sanitation programme has been completed and the number of microorganisms remaining as assessed by traditional microbiological techniques, is not likely to be useful. They are cheaper in use than ATP based systems as the end point of the tests is a visible colour change rather than a signal which is interpreted by an instrument e.g. light output measured by a luminometer. 418 Chilled foods Protein hygiene tests were further developed and recently reintroduced by Konica as the ‘Swab’ n’ Check’ hygiene monitoring kit. This assay detects the presence of protein on a surface by an enhanced Biuret reaction, the end point of which is a colour change from green through to purple. The surface to be assessed is swabbed and the swab placed into a tube of resuspension fluid containing the reagents necessary to activate the Biuret reaction. After ten minutes any colour change is compared to a supplied colour card and the degree of colour change used as an indication of the hygienic nature of the surface. However, there is currently little published data on both the efficacy of this system (Griffith et al. 1997) and the food-processing environments to which it is best suited. Other manufacturers have also recently launched competing products. Verification of the performance of the sanitation programmes is usually undertaken by microbiological methods in the chilled food industry, though ATP levels are also used (especially in low risk). Microbiological sampling is typically for the total number of viable microorganisms remaining after cleaning and disinfection, i.e. total viable count (TVC), both as a measurement of the ability of the sanitation programme to control all microorganisms and to maximise microbial detection. Sampling targeted at specific pathogens or spoilage organisms, which are thought to play a major role in the safety or quality of the product, is undertaken to verify the performance of the sanitation programme designed for their control. Microbiological assessments have also been used to ensure compliance with external microbial standards, as a basis for cleaning operatives’ bonus payments, in hygiene inspection and troubleshooting exercises, and to optimise sanitation procedures. Traditional microbiological techniques appropriate for food factory use involving the removal or sampling of microorganisms from surfaces, and their culture using standard agar plating methods have been reviewed by Holah (1998). Microorganisms may be sampled via sterile cotton or alginate swabs and sponges, after which the microorganisms are resuspended by vortex mixing or dissolution into suitable recovery or transport media, or via water rinses for larger enclosed areas (e.g. fillers). Representative dilutions are then incubated in a range of microbial growth media, depending on which microorganisms are being selected for, and incubated for 24–48 hours. Alternatively, microorgan- isms may be sampled directly onto self-prepared or commercial (‘dip slides’) agar contact plates. The choice of sampling site will relate to risk assessment. Where there is the potential for microorganisms remaining after (poor) cleaning and disinfection to, via e.g. direct product contact, infect large quantities of product, these sources would require sampling much more frequently than other sites which, whilst they may be more likely to be contaminated, pose less of a direct risk to the product. For example, it is more sensible, and gives more confidence, to sample the points of the equipment that directly contact the product and that are difficult to clean than to sample non-direct contact surfaces, e.g. underneath of the equipment framework. Cleaning and disinfection 419 In relation to microorganism numbers, it is difficult to suggest what is an ‘acceptable’ number of microorganisms remaining on a surface after cleaning and disinfection as this is clearly dependent on the food product, process, ‘risk area’ and degree of sanitation undertaken. A number of figures have been quoted in the past (as total viable count per square decimeter) including 100 (Favero et al. 1984), 540 (Thorpe and Barker 1987) and 1000 (Timperley and Lawson 1980) for dairies, canneries and general manufacturing respectively. The results in Table 14.1 show that in chilled food production, sanitation programmes should achieve levels of around 1000 microorganisms per swab, which on flat surfaces approximately equates to a square decimeter. Expressing counts arithmetically is always a problem, however, as single counts taken in areas where cleaning has been inadequate (which may be in excess of 10 8 per swab) produce an artificially high mean count, even over thousands of samples. It is better, therefore, to express counts as log to the base 10, a technique that places less emphasis on a relatively few high counts, and Table 14.1 shows that log counts of approximately 1 should be obtained. Because of the difficulty in setting external standards, it is best to set internal standards as a measurement of what can be achieved by a given sanitation programme. A typical approach would be to assess the level of microorganisms or ATP present on a surface after a series of ten or so sanitation programmes in which the sanitation programme is carefully controlled (i.e. detergent and disinfectant concentrations are correct, contact times are adhered to, water temperatures are checked, pressure hoses are set to specified pressures, sanitation schedules are followed, etc.). The mean result will provide an achievable standard (or standards if specific areas differ significantly in their cleanability) which can be immediately used and can be reviewed as subsequent data points are obtained in the future. A review of the standard would be required if either the food product or process or the sanitation programme were changed. As part of the assessment of sanitation programmes, it is worthwhile looking how the programme is performing over a defined time period (weekly, monthly, quarterly etc.) as individual sample results are only an estimate of what is happening at one specific time period. This may be to ensure that the programme remains within control, to reduce the variation within the programme or, as should be encouraged, to try and improve the programme’s performance. An assessment of the performance of the programme with time, or trend analysis, can be undertaken simply, by producing a graphical representation of the results on a time basis, or can be undertaken from a statistical perspective using Statistical Process Control (SPC) techniques as described by Harris and Richardson (1996). Generally, graphical representation is the most widely used approach, though SPC techniques should be encouraged for more rigorous assessment of improvement in the programme’s performance. 420 Chilled foods 14.7 Management responsibilities Senior management must take full responsibility for the successful operation of the sanitation programme; ultimately, failures in the programme generally reflect poor management. For the majority of chilled food processing operations, the following is a guide to the responsibilities of senior management. ? Always seek to improve hygiene standards in line with the high-risk philosophies adopted in Chapter 13. Hygiene has traditionally not had the same research support as other areas of importance in food manufacture and is thus a new and developing science. It is only relatively recently that new concepts have been developed, based on scientific assessments, and management must be flexible enough to try out and to encourage such concepts when they emerge. ? Lead by example by being both always properly attired in food production areas and (occasionally) present in production areas when sanitation is being undertaken (usually in the early hours of the morning!) ? Provide the required equipment (including maintenance), the staffing levels and the time to undertake the sanitation programme effectively. Cleaning operatives must be a dedicated labour pool whose priority is to sanitation (i.e. not production). Similarly, operatives should not join in the cleaning team as an ‘introduction to production’. ? Management should be capable of giving praise when sanitation is undertaken correctly, as well as discipline when it is not. In companies where bonus systems have been employed based on microbiological assessments of equipment after cleaning, results have indicated that hygiene has generally been improved and bonuses are rarely missed. ? Appoint or nominate a manager to be responsible for the day-to-day implementation of the sanitation programme. The manager who assumes responsibility for the sanitation programme must have technical hygiene expertise and has a range of job functions including the following: ? the selection of a suitable chemical supplier ? the selection of sanitation chemicals, equipment and methodology ? the training of cleaning operatives ? the development of cleaning schedules ? the implementation of sanitation programme monitoring systems ? representation of hygiene issues to senior management. Good chemical suppliers are able to do much more than simply supply detergents and disinfectants. They should be chosen on their abilities to undertake site hygiene audits, supply suitable chemical dosing and application equipment, undertake operative training and help with the development of cleaning schedules and sanitation monitoring and verification systems. Good chemical companies respond quickly to their customers’ needs, periodically Cleaning and disinfection 421 review their customers’ requirements and visit during sanitation periods to ensure that their products are being used properly and are working satisfactorily. The cleaning manager may also need to visit the chemical supplier’s site to audit their quality systems. Whilst in theory systems and/or chemicals could seem appropriate for the required task, every factory, with its water supply, food products, equipment, materials of construction and layout, etc., is unique. All sanitation chemicals, equipment and methodology must, therefore, be proven in the processing environment. New products and equipment are always being produced and a good working relationship with hygiene suppliers is beneficial. Only disinfectants that have been approved to the relevant European Standards should be used. The cleaning operatives’ job is both technical and potentially hazardous, and all steps should be undertaken to ensure that sufficient training is given. By the nature of the job, training is likely to be comprehensive and should include: ? a knowledge of basic food hygiene ? the importance of maintaining low/high-risk barriers during cleaning ? the implications to product safety/spoilage of poor sanitation practices ? an understanding of the basic function and use of sanitation chemicals and equipment and of their sequence of operation ? a thorough knowledge of the safe handling of chemicals and their application and the safe use of sanitation equipment. For each piece of equipment or for each processing area, a cleaning or sanitation schedule should be developed, preferably in a loose-leaf format so that it can readily be updated and which should always be available for inspection by cleaning operatives or auditors. The schedule must show clearly each stage of the cleaning and disinfection process (diagrammatically if this would help), all pertinent information on safety, and the key inspection points and how these should be assessed. It is difficult to produce a list of requirements that should be found in a cleaning schedule, but the following is a typical, non-exhaustive list: ? a description, hazard code, in-use concentration, method of make up, storage conditions, location and amount to be drawn of all chemicals used ? type, use of, set parameters (pressure, nozzle type, etc.), maintenance and location of sanitation equipment ? description of the equipment to be cleaned, need for fitters, need to disconnect from services, dismantling and reassembly procedures ? full description of the cleaning process, its frequency and requirement for periodic measures ? staff requirements and their responsibilities ? key points for assessment of the sanitation procedure and description of evaluation procedures for programme monitoring and verification. When new equipment is purchased or processing areas designed or refurbished, insufficient attention is usually placed on sanitation requirements. 422 Chilled foods Equipment or areas of poor hygienic design will be more expensive to clean (and maintain) and may not be capable of being cleaned to an acceptable standard in the time available. If improperly cleaned, adequate disinfection is impossible and thus contamination will not be controlled. Hygiene management must be strongly represented, thus ensuring that hygiene requirements are considered alongside those of engineering, production and accounts, etc. Three types of sanitation programme can be implemented by management and each has its advantages and disadvantages: at the end of production, production operatives clean their workstations and then (a), they form a cleaning crew and undertake the sanitation programme; (b), a separate, dedicated cleaning gang complete the sanitation programme; or (c) cleaning and disinfection is undertaken by contract cleaners. Whilst each option will place different demands on the food manufacturer, the principles as mentioned above should always be incorporated and the sanitation programme effectively managed. 14.8 References ANON, (1964) Assessment of odour from packaging material used for foodstuffs. British Standard 3744. British Standards Institute, London. ANON, (1973) The effectiveness of water blast cleaning in the food industry. Food Technology in New Zealand 8, 15 and 21. ANON, (1983a) Sensory analysis of food. Part 5: Triangular test. ISO 4120. International Standards Organisation. ANON, (1983b) Testing of container materials and containers for food products. DIN 10955. Deutsches Institut fu¨r Normung e.V., Berlin. ANON, (1991) Sanitation. In: Shapton, D.A. and N.F. (eds) Principles and Practices for the Safe Processing of Foods. Butterworth-Heinemann Ltd., Oxford, pp. 117–99. ANON, (1997) EN 1276:1997 Chemical disinfectants and antiseptics – Qunatitative suspension test for the evaluation of bactericidal activity of chemical disinfectants and antiseptics used in food, industrial, domestic, and institutional areas – Test method and requirements (phase 2, step 1). ANON, (1998a) EN 1650:1998 Chemical disinfectants and antiseptics – Quantitative suspension test for the evaluation of fungicidal activity of chemical disinfectants and antiseptics used in food, industrial, domestic, and institutional areas – Test method and requirements (phase 2, step 1). ANON, (1998b) A practical guide to the disinfection of food processing factories and equipment using fogging. Ministry of Agriculture Fisheries and Food, 650, St. Catherine Street, London, SE1 0UD. ANDERSON M E, HUFF H E and MARSHALL R T, (1985) Removal of animal fat from food grade belting as affected by pressure and temperature of sprayed water, Journal of Food Protection, 44 246–8. ANDRADE N J, BRIDGEMAN T A and ZOTTOLA E A, (1998) Bacteriocidal activity of Cleaning and disinfection 423 sanitizers against Enterococcus faecium attached to stainless steel as determined by plate count and impedance methods. Journal of Food Protection, 661, 833–8. BAUTISTA D A, MCINTYRE L, LALEYE L and GRIFITHS M W, (1992) The application of ATP bioluminescence for the assessment of milk quality and factory hygiene, Journal Rapid Methods Autom Microbiology, 1 179–93. BELL C, STALLARD P A, BROWN S E and STANDLEY J T E, (1994) ATP- Bioluminescence techniques for assessing the hygienic condition of milk transport tankers, International Dairy Journal, 4 629–40. BERGMAN B-O and TRAGARDH C, (1990) An approach to study and model the hydrodynamic cleaning effect, Journal of Food Process Engineering, 13 135–54. BLENKINSOPP S A and COSTERTON J W, (1991) Understanding bacterial biofilms, Trends in Biotechnology, 9 138–43. CARPENTIER B and CERF O, (1993) Biofilms and their consequences, with particular reference to the food industry, Journal of Applied Bacteriology, 75 499–511. CORRIEU G, (1981) State-of-the-art of cleaning surfaces. In Proceedings of Fundamentals and Application of Surface Phenomena Associated with Fouling and Cleaning in Food Processing, April 6–9, Lund University, Sweden, pp. 90–114. DAS J R, BHAKOO M, JONES M V and GILBERT P, (1998) Changes in biocide susceptibility of Staphylococcus epidermidis and Escherichia coli cells associated with rapid attachment to plastic surfaces, Journal of Applied Microbiology, 84 852–8. DESTRO M T, LEITAO M F F and FARBER J M, (1996) Use of molecular typing methods to trace the dissemination of Listeria monocytogenes in a shrimp processing plant, Applied and Environmental Microbiology, 62 705–11. DHALIWAL D S, CORDIER J L and COX L J, (1992) Impedimetric evaluation of the efficacy of disinfectants against biofilms, Letters in Applied Microbiology, 15 217–21. DUNSMORE D G, (1981) Bacteriological control of food equipment surfaces by cleaning systems. 1. Detergent effects, Journal of Food Protection, 44 15– 20. DUNSMORE D G, TWOMEY A, WHITTLESTONE W G and MORGAN H W, (1981) Design and performance of systems for cleaning product contact surfaces of food equipment: A review, Journal of Food Production, 44 220–40. ELLIOT R P, (1980) Cleaning and sanitation. In Katsuayama, A.M. (ed.) Principles of food processing sanitation, The Food Processors Institute, USA, pp. 91–129. FAVERO M S, GABIS D A and VESELEY D, (1984) Environmental monitoring procedures. In: Speck, M. L. (ed.) Compendium of Methods for the Microbiological Examination of Foods, American Public Health Associa- tion, Washington, DC. FRANK J F and KOFFI R A, (1990) Surface-adherent growth of Listeria 424 Chilled foods monocytogenes is associated with increased resistance to surfactant sanitizers and heat, Journal of Food Protection, 53 550–4. GIBSON H, TAYLOR J H, HALL K E and HOLAH J T, (1995) Biofilms and their detection in the food industry, CCFRA R&D Report No. 1, CCFRA, Chipping Campden, Glos., GL55 6LD, UK. GIBSON H, TAYLOR J H, HALL K E and HOLAH J T, (1999) Effectiveness of cleaning techniques used in the food industry in terms of the removal of bacterial biofilms. Journal of Applied Microbiology, 87 41–8. GRIFFITH C J, DAVIDSON C A, PETERS A C and FIELDING L M, (1997) Towards a strategic cleaning assessment programme: hygiene monitoring and ATP luminometry, an options appraisal, Food Science and Technology Today, 11 15–24. GRIFFITHS J, BLUCHER A, FLERI J and FIELDING L, (1994) An evaluation of luminometry as a technique in food microbiology and a comparison of six commercially available luminometers, Food Science and Technology Today, 8 209–16. HARRIS C S M and RICHARDSON P S, (1996) Review of Principles, Techniques and Benefits of Statistical Process Control. Review No. 4, Campden and Chorleywood Food Research Association, Chipping Campden, UK. HAWRONSKYI J-M and HOLAH J T, (1997) ATP: A universal hygiene monitor, Trends in Food Science and Technology, 8 79–84 HAYES P R, (1985) Cleaning and Disinfection: Methods. In: Food Microbiology and Hygiene, Elsevier Applied Science Publishers, Barking, pp. 268–305. HOLAH J T, (1991) Food Surface Sanitation. In: Hui, Y H (ed.) Encyclopedia of Food Science and Technology, John Wiley and Sons, New York. HOLAH J T, (1995) Special needs for disinfectants in food-handling establish- ments, Revue Scientific et Technique Office International des E ′ pizooties, 14 95–104. HOLAH J T, (1998) Guidelines for Effective Environmental Microbiological Sampling, Guideline No. 20, Campden and Chorleywood Food Research Association, Chipping Campden, UK. HOLAH J T and KEARNEY L R, (1992) Introduction to biofilms in the food industry. In Biofilms – Science and Technology, edited by L.F. Melo, T.R. Bott, M. Fletcher and B. Capdeville, pp. 35–45. Dordrecht: Kluwer. HOLAH J T and THORPE R H, (1990) Cleanability in relation to bacterial retention on unused and abraded domestic sink materials, Journal of Applied Bacteriology, 69 599–608. HOLAH J T, HIGGS C, ROBINSON S, WORTHINGTON D and SPENCELEY H, (1990a) A Malthus based surface disinfection test for food hygiene, Letters in Applied Microbiology, 11 255–9. HOLAH, J T, LAVAUD A, PETERS W and DYE K A, (1998) Future techniques for disinfectant testing, International Biodeterioration and Biodegradation, 41 273–9. HOLAH J T, TIMPERLEY A W and HOLDER J S, (1990b) The spread of Listeria by cleaning systems. Technical Memorandum No. 590, Campden Food and Cleaning and disinfection 425 Drink Research Association, Chipping Campden, UK. HUGO W B, PALLENT L J, GRANT D J W, DENYER S P and DAVIES A, (1985) Factors contributing to the survival of Pseudomonas cepacia in chlorhexidine, Letters in Applied Microbiology, 2 37–42. HUGO W B and RUSSELL A D, (1999) Types of antimicrobial agent. In Principles and Practices of Disinfection, Preservation and Sterilization. Eds A.D. Russell, W.B. Hugo and G.A.J. Ayliffe, Blackwell Sciences, Oxford, pp. 5– 95. ICMSF, (1980) Cleaning, disinfection and hygiene. In Microbial Ecology of Foods, Volume 1, Factors Affecting Life and Death of Microorganisms. The International Commission on Microbiological Specifications for Foods, Academic Press, London, pp. 232–58. ICMSF, (1988) Cleaning and disinfecting. In Microorganisms in foods, Volume 4, Application of the Hazard Analysis Critical Control Point (HACCP) System to Ensure Microbiological Safety and Quality. The International Commis- sion on Microbiological Specifications for Foods, Blackwell Scientific, Oxford, pp. 93–116. JENNINGS W G, (1965) Theory and practice of hard-surface cleaning, Advances in Food Research, 14 325–459. KLEMPERER R, (1982) Tests for disinfectants: principles and problems. In: Disinfectants: their assessment and industrial use, Scientific Symposia Ltd, London. KNAUER-KRAETZL B, (1994) Reinigung und Desinfektion. 12. Fleischwarenfor- um: Fleischhygiene, BEHR’s Seminar, Darmstadt, 14–15 April 1994. KOOPAL L K, (1985) Physico-chemical aspects of hard-surface cleaning, Netherlands Milk and Dairy Journal, 39 127–54. KUMAR C G and ANAND S K, (1998) Significance of microbial biofilms in the food industry: a review, International Journal of Food Microbiology, 42 9–27. LAWRENCE L M and GILMORE A, (1995) Characterisation of Listeria mono- cytogenes isolated from poultry products and poultry-processing environ- ment by random amplification of polymorphic DNA and multilocus enzyme electrophoresis, Applied and Environmental Microbiology, 61 2139–44. LE CHEVALIER M W, CAWTHORN C D and LEE R G, (1988) Inactivation of biofilm bacteria, Applied and Environmental Microbiology, 44 972–87. LEE S-L and FRANK J F, (1991) Effect of growth temperature and media on inactivation of Listeria monocytogenes by chlorine, Journal of Food Safety, 11 65–71. LONCIN M, (1977) Modelling in cleaning, disinfection and rinsing. In: Proceedings of Mathematical modelling in food processing,7–9 September, Lund. Lund Institute of Technology. MABESA R C, CASTILLO M M, CONTRERAS E A, BANNAAD L and BANDIAN V, (1982) Destruction and removal of microorganisms from food equipment and utensil surfaces by detergents, The Philippine Journal of Science, 3 17–22. MARRIOT N G, (1985) Sanitation Equipment and Systems. In Principles of Food Sanitation, AVI Publishing Co., Westport, CT, pp. 117–51. 426 Chilled foods MATTILA-SANDHOLM T and WIRTANEN G, (1992) Biofilm formation in the food industry: a review, Food Reviews International, 8, 573–603. METTLER E and CARPENTIER B, (1998) Variations over time of microbial load and physicochemical properties of floor materials after cleaning in food industry premises, Journal of Food Protection, 61 57–65. MIDDLEMISS N E, NUNES C A, SORENSEN J E and PAQUETTE G, (1985) Effect of a water rinse and a detergent wash on milkfat and milk protein soils, Journal of Food Protection, 48 257–60. MROZEK H, (1982) Development trends with disinfection in the food industry, Deutsche-Molkerei-Zeitung, 12 348–52. OFFILER M T, (1990) Open plant cleaning: equipment and methods. In Proceedings of C96Hygiene for the 90s’, November 7–8, Campden Food and Drink Research Association, Chipping Campden, Glos., UK, pp. 55–63. PONTEFRACT R D, (1991) Bacterial adherence: its consequesnces in food processing, Canadian Institute of Food Science and Technology Journal, 24 113–17. POULIS J A, DE PIJPER M, MOSSEL D A A and DEKKERS P P H A, (1993). Assessment of cleaning and disinfection in the food industry with the rapid ATP- bioluminescence technique combined with tissue fluid contamination test and a conventional microbiological method, International Journal of Food Microbiology, 20 109–16. REYBROUCK G, (1998) The testing of disinfectants, International Biodeteriora- tion and Biodegradation, 41 269–72. RIDGEWAY H F and OLSEN B H, (1982) Chlorine resistance patterns of bacteria from two drinking water distribution systems, Applied and Environmental Microbiolog, 44 972–87. RONNER A B and WONG A C L, (1993) Biofilm development and sanitizer inactivation of Listeria monocytogenes and Salmonella typhimurium on stainless steel and Buna-N rubber, Journal of Food Protection, 56 750–8. RUSSELL A D, HUGO W B and AYLIFFE G A J, (1982) Principles and Practice of Disinfection, Preservation and Sterilization, Blackwell Scientific Publica- tions, London. SCHLUSSLER H J, (1975) Zur Kinetic von Reinigungsvorgangen an festen Oberflachen. Symposium u¨ber Reinigen und Desinfizieren lebensmittelver- arbeitender Anlagen, Karlsruhe, 1975. SCHMIDT U and CREMLING K, (1981) Cleaning and disinfection processes. IV. Effects of cleaning and other measures on surface bacterial flora, Fleischwirtschaft, 61 1202–7. SHUPE W L, BAILEY J S, WHITEHEAD W K and THOMPSON J E, (1982) Cleaning poultry fat from stainless steel flat plates, Transactions of the American Society of Agricultural Engineers, 25 1446–9. TAYLOR J H, ROGERS S J and HOLAH J T, (1999) A comparison of the bactericidal efficacy of 18 disinfectants used in the food industry against Escherichia coli 0157:H7 and Pseudomonas aeruginosa at 10oC and 20oC, Journal of Applied Microbiology, 87 718–25. Cleaning and disinfection 427 THORPE R H and BARKER P M, (1987) Hygienic Design of Liquid Handling Equipment for the Food Industry. Technical Manual No. 17, Campden Food and Drink Research Association, Chipping Campden, UK. TIMPERLEY D A and LAWSON G B, (1980) Test rigs for evaluation of hygiene in plant design. In Jowitt, R. (ed.), Hygienic Design and Operation of Food Plant. Ellis Horwood, Chichester. WALKER R L, JENSEN L H, KINDE H, ALEXANDER A V and OWENS L S, (1991) Environmental survey for Listeria species in frozen milk product plants in California, Journal of Food Protection, 54, 178–82. WIRTANEN G and MATTILA-SANDHOLM T, (1993) Epifluorescence image analysis and cultivation of foodborne bacteria grown on stainless steel surfaces, Journal of Food Protection, 56 678–83. WIRTANEN G and MATTILA-SANDHOLM T, (1994) Measurement of biofilm of Pediococcus pentosacceus and Pseudomonas fragi on stainless steel surfaces, Colloids and Surfaces B: Biointerfaces, 2 33–9. WRIGHT J B, RUSESKA I and COSTERTON J W, (1991) Decreased biocide susceptibility of adherent Legionella pneumophila, Journal of Applied Microbiology, 71 531–8. ZOTTOLA E A and SASAHARA K C, (1994) Microbial biofilms in the food processing environment – should they be a concern? International Journal of Food Microbiology, 23 125–48. 428 Chilled foods