Sei sulla pagina 1di 10

food and bioproducts processing 8 8 ( 2 0 1 0 ) 365374

Contents lists available at ScienceDirect

Food and Bioproducts Processing


journal homepage: www.elsevier.com/locate/fbp

Characterising the cleaning mechanisms of yeast and the implications for Cleaning In Place (CIP)
K.R. Goode a,b , K. Asteriadou a , P.J. Fryer a, , M. Picksley b , P.T. Robbins a
a b

School of Chemical Engineering, University of Birmingham, Edgbaston, Birmingham B15 2TT, UK Heineken UK, 24 Broadway Park, South Gyle Edinburgh EH12 9JZ, UK

a b s t r a c t
Deposition of yeast inside brewery process plant is a serious industrial problem. Investigation of the cleaning of beer fermenter deposits revealed two types of fouling; yeast foam (type A) and yeast lm (type B). Rheological characterisation indicated both deposits could be mimicked in lab scale fouling experiments using yeast slurry aged for different times. Water and chemical rinsing of these deposits on a lab scale ow cell revealed three distinct cleaning phases: (i) hydration and swelling, (ii) removal in the ow by dissolution and in patches and (iii) no further removal. At 30 and 50 C water rinsing at the ow velocities investigated could remove up to 85% of the deposit. At a water rinsing temperature of 70 C, less deposit could be removed overall. Rheological studies indicated that increasing the temperature of the deposit generated a more elastic deposit which may decrease cleanability. Chemical cleaning using 2 wt% Advantis 210 (a NaOH base cleaning agent) eventually gave a visually clean surface at all ow velocities and temperatures. Chemical cleaning at 70 C gave the shortest cleaning times for all ow velocities, but comparable cleaning times were observed when rinsing at 30 and 50 C, suggesting that an increase in temperature from 30 to 50 C might not decrease the cleaning time. 2010 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved. Keywords: Yeast; Fermentation fouling; Water rinsing; Alkali chemical rinsing

1.
1.1.

Introduction
Cleaning yeast in brewery operations

Cleaning is required to avoid microbial contamination in food and beverage manufacture. Automated Cleaning In Place (CIP) is the ubiquitous process used to return the plant to a clean state (Tamime, 2008). It can be argued that if fouling did not occur there would be little need for cleaning, but no economic fouling prevention method has yet been demonstrated, thus research into efcient cleaning remains very important. There are numerous operations involved in making beer, illustrated in Fig. 1. Each stage has a level of cleanliness that needs to be achieved and fouling is encountered at each stage. Fermentation tanks must be both microbiologically clean to avoid contamination of the following batch and rinsed completely of cleaning chemical to avoid product contamination (Salo et al., 2008). A typical fermenter CIP regime carried out at a brew-

ery studied in this project is listed in Table 1. A ow velocity of at least 1.5 m s1 is used. In a CIP sequence the pre-rinse and chemical phases of cleaning have the most impact on the amount of material removed. In a brewery, yeast is used to ferment sugar extracted from malt to make beer. This is carried out in conical fermentation vessels that hold up to 12,000 hectolitres in large scale operations. Fouling deposits have been observed (Cluett, 2001) here classied for ease of referral as: Type A. Formed during fermentation above the beer level. This deposit can age on the surface for up to 7 days. Type B. Residual yeast attached to the vessel wall and cone during emptying; this deposit can age on the surface for up to 5 h. Various authors have examined yeast removal from surfaces. Yeast readily attaches to stainless steel and plastics

Corresponding author. Tel.: +44 (0)121 414 5451; fax: +44 (0)121 414 5324. E-mail address: p.j.fryer@bham.ac.uk (P.J. Fryer). Received 11 May 2010; Received in revised form 13 August 2010; Accepted 18 August 2010 0960-3085/$ see front matter 2010 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved. doi:10.1016/j.fbp.2010.08.005

366

food and bioproducts processing 8 8 ( 2 0 1 0 ) 365374

Nomenclature Abbreviations CAM cellulose acetate membrane CIP Cleaning In Place EHEDG European Hygienic Engineering & Design Group LVR linear viscoelastic region Symbols C cf de G G Lw My q Re Tav Tc U V v
a

w50%

consistency of yeast slurry (wt%) friction factor equivalent diameter (m) elastic modulus (N m2 (Pa)) viscous modulus (N m2 (Pa)) volume of wort (L) mass yeast slurry (g) heat ux (kW m2 ) Reynolds number (dimensionless) average temperature of the test section ow ( C) thermocouple ( C) heat transfer coefcient (kW m2 K1 ) viability of yeast slurry (%) ow velocity (m s1 ) shear stress (s1 ) apparent dynamic viscosity (Ns m2 ) density (kg m3 ) the concentration of yeast slurry added to a fermenter (g L1 ) wall shear stress required to remove 50% of attached cells from a surface (N m2 ) wall shear stress (N m2 )

Fig. 1 Schematic of brewery operations (left) and the fouling types encountered at each stage (middle). The cleaning requirements at each stage are also indicated (right). The closer the process stage to nal product the more stringent the level of cleaning becomes.

(Guillemot et al., 2006), elastomers (Chandra et al., 2001) and glass (Mercier-Bonin et al., 2004) all of which are materials using extensively in the beer brewing and dispense industries. Guillemot et al. (2006) investigated the detachment of yeast in a ow chamber at a pH of 5.5, yeast cells are negatively charged over the pH range 3.57. It was found that yeast cells could be wholly removed from glass using water, and that yeast had a strong adhesion to stainless steel. The wall shear stress required to remove 50% of the attached cells from stainless steel, denoted w50% was 30 Pa, whilst for plastics w50% ranged from 1 to 2 Pa. Mozes et al. (1987) found that yeast could attach and form a dense layer of cells on stainless steel and aluminium at pH 3 and 56, and that a dense layer of yeast cells would attach to glass and plastics if the negative charge was reduced by treatment with ferric ions. Deposits

formed by reaction processes or microbes usually cannot be wholly removed with water from stainless steel (Christian et al., 2006). During the microltration of beer, yeast readily fouls the membranes. Gell et al. (1999) found that increasing the concentration of yeast in solution with the protein resulted in lower uxes and intermediate protein transmission through the cellulose acetate membrane (CAM). When the yeast cells were present on the membrane as a layer, termed a yeast cake, the yeast was believed to form a second membrane on the CAM. Increasing the thickness of this layer reduced the permeate ux and protein transmission. Mores and Davis (2002) examined the effect of pulsing ow through a CAM. They found that ux increased with increasing shear rate, backpulse pressure and backpulse duration. At higher shear rate and backpulse pressure multiple short backpulses were more effective in cleaning the membrane. At low shear rate and backpulse pressure fewer, longer backpulses were more effective. Longer, weaker backpulses led to the highest recovered uxes. Increasing the ow rate or temperature of water rinses has been shown to aid deposit removal (Friis and Jensen, 2002), and ow pulsing has been shown to affect milk protein cleaning (Gillham et al., 2000).

1.2.

Approach adopted

Until recent years the efciency and impact of cleaning has been largely neglected, as:

Table 1 Frequently used fermenter CIP regime. Stage


1 2 3 4 5 Pre-rinse Chemical recirculation Intermediate rinse Sterilant Post-sterilant rinse

Purpose
Remove bulk soil by dissolution and/or shear force to leave the surface less fouled. Circulation of alkali-based chemicals to dissolve and remove remaining material and kill microbes. Circulation of water/chemical to neutralise pH and remove traces of remaining chemical and material. Circulation of chemical/steam around the clean plant to kill remaining microbes. If necessary, sterile water rinse to remove remaining traces of the chemical.

Conditions
Ambient, 10 min 70 C, 0.12 wt%, 20 min Ambient, 5 min Steam or ambient chemical, 15 min Ambient, 5 min

Adapted from ORourke (2003).

food and bioproducts processing 8 8 ( 2 0 1 0 ) 365374

367

A signicant body of knowledge on cleaning exists within individual manufacturers, cleaning chemical companies, and organisations such as the European Hygienic Engineering & Design Group (EHEDG), that has produced extensive guidelines on the types of surface and equipment that are easy to clean (EHEDG Yearbook, 2009). Compartmentalisation of cleaning information has resulted in independent development of a good way to clean a product. Comparison of cleaning information for different products and scale up of cleaning to industrial scale equipment is thus difcult and semi-empirical. Fryer and Asteriadou (2009) suggested a classication of cleaning problems in terms of cleaning cost and soil complexity, with three deposit types that represent a range of different rheological and cleaning behaviours: type 1 deposit: a viscoelastic or viscoplastic uid (such as toothpaste) that can be rinsed with water alone, type 2 deposit: a biological lm removed in part by water and in part by chemical, and type 3 deposit: a hard cohesive deposit that cannot be removed by water alone and that requires chemical action for removal. These three cleaning regimes have been investigated in different parts of the ZEAL programme (see Cole et al., 2010; Othman et al., 2010; Akhtar et al., 2010; Sahu et al., 2007). Rheological characterisation of type 1 deposits such as yoghurt has permitted removal during water rinsing from a pipe to be modelled (Henningsson et al., 2007). Knowledge of foulant material behaviour under different ow conditions is thus important in determining how it can be removed from a surface. Here a study of yeast deposits is made to identify both their rheological and cleaning behaviour.

Fig. 2 Factors that contribute to the effects of beer fermentation CIP in terms of (a) cost and (b) tonnes of CO2 . CIP is a non-added value process, commonly considered separately from production, whose true cost is unknown. The cost of brand damage from a product recall is substantial, resulting in little incentive to experiment with CIP parameters such as cleaning time: If it isnt broke dont x it. There are however numerous drivers for a revision of CIP operations including the need to minimise utility usage (energy and water), minimisation of waste and green house gas (GHG) emissions, and the need for product safety and quality. In addition, the increasing use of real time microbiology enables rapid measurement of contamination. Adjustment of process variables (temperature, time, chemical species and ow velocity) to minimise cleaning costs is increasingly important, such as moving to lower temperatures and cleaning chemical concentrations. This project has aimed to improve the efciency of fermentation CIP by: Identifying CIP resource usages, costs, and CO2 emissions by benchmarking. Characterising the cleaning mechanisms. Measuring the cleaning phenomena by in-line measurement techniques. A benchmark of CIP performance was carried out on brewery fermentation vessels. This study was designed with Prof. Roger Benson based on his manufacturing benchmark tools (Benson and McCabe, 2004). The factors that contributed to cost and GHG emissions in this CIP operation were: yield loss, water, efuent, steam, electricity, caustic, Stabilon WT (caustic additive), and P3 Oxysan ZS (sanitizer). Fig. 2 indicates the percentage contribution to (a) cost and (b) GHG emissions. The study gave quantitative data that indicated the use of heated caustic was the biggest contributor to cost and environmental impact, so reducing temperature and caustic concentration would be of value.

2.

Methods and materials

The experimental rig used here is similar to that used previously to characterise cleaning behaviour of various type 1 and type 3 deposits including whey protein and tomato paste (Christian et al., 2006), egg albumin (Liu et al., 2007), toothpaste (Cole et al., 2010), and sweetened condensed milk (Othman et al., 2010). Cleaning can be quantied using image analysis and heat ux measurement. This bench top ow cell rig has been used here to study the removal behaviour of a type 2 soil, namely yeast deposits. The type A deposit (yeast foam) can be cleaned partially by water rinsing and completely with chemical cleaning. Here, experiments have studied the effects of chemical concentration, ow velocity and temperature on cleaning times. The type B deposit (residual yeast) can be removed by water rinsing.

2.1.

Forming the deposits for cleaning

Stainless steel surfaces collected from an industrial fermenter at the end of fermentation had variable levels of fouling. Some surfaces were only half fouled. Yeast slurry, i.e. yeast recovered from fermentation was used as a fouling deposit. This ensured that each test surface had 100% deposit coverage for each cleaning experiment. Deposits were formed on square stainless steel coupons (AISI 316) 2 mm thick, 30 mm by 30 mm with a milled ridge 2.5 mm 1 mm for ease of positioning in the test section. This gives a surface area of 625 mm2 that can be fouled. 1 ml of yeast slurry recovered from fermentation (John Smiths Brewery, Tadcaster) was applied to clean coupons

368

food and bioproducts processing 8 8 ( 2 0 1 0 ) 365374

Fig. 3 Schematic of the ow cell. The microfoil heat ux sensor (MHFS), and thermocouples Tc2 and Tc3 are attached to a copper stub placed directly underneath the fouled coupon. The stub is supported by a spring in a copper block that sits in an ice bath. V manual valve, C conductivity probe, Tc thermocouple. using a BioHit pipette (BioHit, Devon) and widened tips. Slurry samples were collected weekly from a yeast storage tank at the brewery, refrigerated and used within 2 days. The tank was selected according to brewery records and agitated prior to sample collection. Slurry samples were taken from storage tanks that had a total suspended solids content of up to 40 wt%. The sample point was rinsed using de-aerated water and soaked in Savlon spray cleaner for 5 min. Cell density and viability was determined prior to fouling using a haemocytometer (Neubauer, Sussex) and light microscope (Olympus BX50, Japan). 1 ml of yeast slurry was added to 9 ml of distilled water, vortexed, and 1 ml immediately extracted and added to 9 ml of water. This serial dilution was repeated once more, 1 ml extracted and added to 1 ml of methylene violet. The mixture was vortexed and 20 l extracted. This was pipetted between the haemocytometer slide and the coverslip to ll the counting chamber. Typically the cell viability was greater than 90% and cell densities were on the order of 1 108 cells per ml. 650 l of slurry per coupon was found to give deposits consistent in appearance, with an average mass of 0.06 0.03 g. To generate type A deposit yeast slurry was incubated on the surfaces at 30 C for 5 days, representing an industrial fermentation time. The percentage of live cells on the surfaces after 5 days aging prior to cleaning experiments was 1%. A portion of deposit from an industrial fermenter was assayed and 34% of the cells were alive. The cell density in the deposit from the fermenter was three orders of magnitude smaller than the yeast slurry. tion. The square duct shaped test section was built to have an equivalent diameter of 25 mm. In cleaning experiments using chemicals, water was pumped from the chemical storage tank to the bypass loop and re-circulated back to the tank to ensure the cleaning uid was of the desired concentration. The chemical used in these experiments was Advantis 210 (Ecolab, Cheadle). The concentration used was 2 wt% Advantis, which contained 1 wt% NaOH and 0.2 wt% KOH. The ow rates achieved through the test section were in the range of 3.617.0 L min1 (0.21.0 m3 h1 ) giving ow velocities in the range of 0.120.6 m s1 . The ow velocities were used to calculate Reynolds numbers (Re) and wall shear stresses according to the Blasius correlation: Re =
w

vde
a

(1)

v2

= cf = 0.079Re0.25

(2)

2.2.

Equipment

where 2 500 < Re < 105 . The ow velocities used were 0.13, 0.26, 0.4, and 0.5 m s1 , at 20, 30, 50, and 70 C, giving Reynolds numbers in the range 286035,000 and surface shear stresses of 0.061.24 N m2 . The ow velocities, Reynolds numbers, and wall shear stresses achieved in the lab scale rig are smaller than those generated in industrial CIP, such as the pilot scale experiments of Cole et al. (2010). As such the smaller rig was used to assess if type A soil can be removed at lower ow rates. Two parameters were evaluated: The area of coupon covered (mm2 ) was determined from images taken using a Canon EOS 30D digital camera (Canon, Japan) attached to a stand (Hama, Germany) positioned over the test section with the coupon in focus. The area was determined for each digital image using software (Image J). The deposit and surface were indistinguishable in grey scale so images were altered in Photoshop CS2 to highlight the

The lab scale cleaning of deposits was assessed using a ow cell similar to that described by Christian (2004) and Aziz (2008). A schematic of the cleaning rig is illustrated in Fig. 3. Water was directed from the mains through a reverse osmosis lter to the water or chemical tank. Water from the storage tank was pumped using a centrifugal pump (Alfa Laval, Denmark) through a coil in the heated tank to the test sec-

food and bioproducts processing 8 8 ( 2 0 1 0 ) 365374

369

Table 2 Characteristics of the fouling deposits formed during fermentation. Fermentation


Industrial scale Pilot scale (miniature fermenter)

Fouling thickness (mm)


3 3

Geometry surface area (m2 )


135.0 0.3

Surface area fouled (m2 )


25.45 0.06

Fouling coverage (%)


18.9 20

deposit. The fouled area was seen to decrease as cleaning progressed. Heat transfer coefcient, U (kW m2 K1 ) was determined according to: U= q Tav Tc2 (3)

from the measured heat ux, q (kW m2 ), the average temperature of the test section cleaning uid (Tav ), and the temperature of Tc2 ( C), the insulated thermocouple positioned below the coupon illustrated in Fig. 3. U increased as cleaning progressed and remained constant when no further deposit was removed, or there appeared to be no deposit remaining on the surface.

carried out in a miniature Cornelius Flask fermenter (The Cornelius Co., MN, USA). Typically the feedstock and the yeast slurry were added to the vessel which is agitated or aerated for a couple of hours. The vessel was then held at around 20 C for up to one week to ferment. This ask could hold up to 15 L of the fermentation feedstock, which here was a sugar (wort) solution extracted from malt and sampled from the outlet of the plate heat exchanger at the brewery (John Smiths Brewery, Tadcaster) when it was en route to a fermentation vessel. The sample tap was sterilised before and after sampling by spraying methylated spirit and aming using a portable bunsen burner (Talentum Development Ltd., Lancashire). The mass of yeast slurry (My ) required for fermentation in the Cornelius Flask in grams was determined from: My = Lw y CV (4)

2.3.

Cleaning rig procedure


where Lw is the volume of wort added to the ask in litres (here 15 L), y is the concentration of yeast slurry added to the industrial fermenter (here 7 g L1 ), C is the consistency, the amount of suspended solids as a percentage of the mixture weight and V is the viability, the fraction of live yeast cells as a percentage of the volume of the mixture. Upon addition of wort and yeast to the Cornelius ask, it was agitated on a stage for a minimum of 2 h before being positioned on a plastic drip tray in a temperature controlled room at 20 C. The ask was then left to ferment for 5 days, representative of an average fermentation time, and the fouling layers were then inspected. In addition, industrial fermenters at the brewery were opened prior to cleaning and the fouling layers observed. Industrial fermenters and the miniature fermentation systems showed a fouling layer above the beer level that looked similar. Both deposits fouled a similar amount of the vessel surface area and were of a similar thickness, summarised in Table 2. This material was sampled and its rheology tested. It was hoped that the rheological characterisation of the deposit generated by Cornelius ask fermentation would resemble that of the deposit generated in the industrial fermenter. Substantial and representative rheological characterisation of the industrial deposit was not feasible but limited tests to assess ow behaviour could be carried out. Deposit from the Cornelius ask was more readily available, easier to handle and could be rheologically characterised more easily.

For all cleaning experiments heat ux data acquisition was started rst. After 5 s the rst image in the series of images was taken using the timer remote controller (TC-80N3, Canon, Japan) attached to the camera. After 10 s the pump was started. For water rinsing experiments the water was re-circulated from the water storage tank. For chemical cleaning experiments water was rst directed from the water tank to the drain for 30 s to prove the route, ensuring the system did not leak. V1 (in Fig. 3) was turned to divert the ow from the chemical tank. Once the conductivity reading stabilised, V6 was closed and V7 opened. At this point chemical was re-circulated through the chemical tank, to minimise chemical use. When a visually clean surface was reached a nal water rinse was done to remove any chemical from the system to make it safe to disassemble in between experiments. Each experiment was ended 300 s after the point when no further deposit was seen to be removed by eye or a visually clean surface was reached. The system was run to this point to ensure no further deposit was removed and U remained constant.

2.4.

Rheological characterisation

The ow properties of yeast soils were determined using an AR500 rheometer (TA Instruments, NJ, USA). The stainless steel parallel plate geometry, 40 mm in diameter, was used to assess the samples. A test gap of 250 m** was set in most cases. Each sample was carefully applied to the centre of the stage using a clean plastic spatula. The sample was left for a minimum of 5 min to equilibrate before testing. A moisture trap was used to keep the water content of the sample constant.

3.
3.1.

Results and discussion


Rheological characterisation

2.5.

Miniature fermentation system

The rheological behaviour of the industrial deposit (at 30 C), miniature fermenter deposit (at 25 C) and yeast slurry (at 30 C) was determined by shear sweeps from 0.01 to 1000 s1 . Fig. 4 shows the apparent viscosity of the three materials over the full range of shear rates tested. The three deposits show: (i) signicantly different behaviour at low shear rates, <1 s1 , with the industrial deposit having a higher viscosity by a

As mentioned in Section 2.1 yeast slurry was used to foul coupons. However the validity of this deposit compared to real fouling needed to be tested. Pilot scale fermentations were

370

food and bioproducts processing 8 8 ( 2 0 1 0 ) 365374

Fig. 4 Viscosity vs. shear rate plot of industrial type A deposit (at 30 C), miniature fermenter deposit (at 25 C), and yeast slurry (at 30 C). factor of 10 than that from the pilot fermenter, and the yeast slurry having a viscosity a factor of ten less; (ii) in the range >2 s1 , the three materials show similar, shear-thinning behaviour. Here the industrial deposit has a lower viscosity. (iii) at higher shear rates, >100 s1 , the three again diverge. From the data it is clear that the industrial deposit has the highest viscosity at low shear, followed by the pilot scale deposit, then the yeast slurry. This may reect the length of time the deposit was aged on the surface which for the industrial deposit was at least 141 h, pilot slurry at least 72 h, and yeast slurry for 1 h. The longer the deposit was aged on the surface the more solid it became. This was also observed by

Mercier-Bonin et al. (2004) who discovered the longer yeast cells were left to contact a glass surface the more strongly the cells were attached. All deposits demonstrated shear-thinning behaviour. The yeast slurry was seen to commence shear-thinning behaviour at a much larger shear rate than the miniature fermenter deposit, around 7 s1 . This was due to the yeast slurry sample retaining liquid. These ndings suggest that higher ow rates would be needed to clean the deposit after a long period of ageing on the surface unless chemical is added to disrupt the deposit structure. Oscillatory stress sweeps of the miniature fermenter deposit and the yeast slurry were conducted at temperatures below 20 C in the range of 0.01 to > 100 Pa, illustrated in Fig. 5(a) and (b). The two sets of curves have similar shapes, with G > G at low stress and G = G at higher stress. The crossover point was found to be at 5.5 Pa for the pilot scale deposit and 1.5 Pa for the yeast slurry. It is notable that the magnitude of the low shear modulus is much greater for the fermenter deposit, at 500 Pa as opposed to 140 Pa at an oscillatory stress of 0.1 Pa. This is most likely due to different testing temperatures; 15 C for miniature fermenter deposit and 18 C for yeast slurry. The maximum surface shear stress achieved in this cleaning system was 1.24 Pa (estimated from Eq. (2)) suggesting the deposits would probably remain predominantly elastic over the ow velocities and temperatures investigated. An oscillatory stress of 0.5 Pa was selected for further investigation of the linear viscoelastic region (LVR) of both deposits. Fig. 5(c) and (d) illustrate that G remained greater than G for both deposits when the temperature was ramped from 20 to 70 C. Increasing the temperature also increased the magni-

Fig. 5 (a) Oscillatory stress sweeps of the miniature fermenter deposit at 15 C, (b) Oscillatory stress sweep of yeast slurry at 18 C, (c) temperature ramp of the miniature fermenter deposit, and (d) temperature ramp of yeast slurry, stress amplitude 0.5 Pa.

food and bioproducts processing 8 8 ( 2 0 1 0 ) 365374

371

tude of the measured modulus, more so for the pilot scale deposit than the yeast slurry. This effect may have been due to the gelation of proteins within the deposits. The difference in the magnitude of the modulus was also expected due to greater moisture retention in yeast slurry. Similar rheological behaviour was shared by the two deposits, indicating that yeast slurry could be used to mimic industrial fouling deposit if the aging time was increased.

3.2.

Deposit removal proles

At the temperatures and ow velocities investigated for water and chemical rinsing, three phases were identied that could be characterised by U and area proles: I. Deposit hydration and swelling; the deposit became lighter in colour. II. Deposit removal by uid ow in part by dissolution and in patches. III. No further removal of deposit; or a visually clean surface was reached. In all cases of water rinsing the surface did not reach a clean state visually; an adhesive lm that could not be removed by further water rinsing remained. A typical coupon is illustrated in Fig. 6(a) showing that the surface is not visually clean. A section of this coupon surface obtained using a surface reectance microscope is shown in Fig. 6(b). The lm appeared to be clusters of yeast cells (in green) on the surface (in yellow). Fig. 7 illustrates U and area proles of deposit removal experiments conducted at a ow velocity of 0.5 m s1 for water and chemical rinsing at ambient and at 70 C. The three phases described above are separated by the vertical dashed lines in each prole. Increasing the temperature decreased both the duration of the swelling phase and the time at which no further deposit removal occurs. The use of 2 wt% Advantis 210 removes all deposit to a visually clean surface, and reduced the duration of the initial swelling phase at all temperatures.

3.3.

The effect of mechanical ow and temperature

Fig. 6 (a) A typical coupon showing that the surface is not visually clean. A section of this coupon surface is shown in (b) on the surface using a surface reectance microscope. The yeast cells are green and the surface is yellow. The deposit was yeast slurry. (For interpretation of the references to colour in this gure legend, the reader is referred to the web version of the article.) data at 70 C is most scattered, there is some evidence that most signicant removal is found at the highest velocity, however there is less removal than at 50 C. Rheological studies indicated that increasing the temperature of the deposit generated a more elastic deposit, illustrated in Fig. 5(b) and (d). This may have resulted in a deposit that was not so susceptible to deformation by the ow. An increase in ow velocity, or the addition of chemical to the ow would be required to remove any more deposit. Water penetration into the deposit whilst rinsing with water appears optimal at 50 C. To remove deposit most quickly and use less water in removing the deposit a rinse temperature of 50 C would be suggested from these results. With a more successful water rinse there would be a reduced deposit load in the chemical cleaning phase, which would help to reduce cleaning chemical use and potentially the temperature of the clean.

At all temperatures (2070 C) and ow velocities (0.260.5 m s1 ), as the rinsing time increased the amount of deposit on the coupon decreased. There was some deposit erosion at the coupon edges and the majority of type A deposit was removed in patches in the ow. The deposit became lighter in colour as rinsing time increased. Fig. 8(a)(c) illustrate the change in the average deposit area with increasing rinsing time at 20, 30, 50, and 70 C. In all cases, initially large patches of deposit were removed, but the size of the removed material decreased with time. Overall the data shows that: Temperature has a bigger impact on deposit removal than ow velocity, in that most of the velocity data shows no trend at each temperature, except at 70 C where there is a clear difference in rinsing at 0.26 and 0.5 m s1 . there is a clear difference between 20 C and the other temperatures, with a lengthy period (ca. 400 s) before removal starts, whilst at other temperatures removal starts more quickly. rinsing at 50 C removed the most deposit in the shortest time.

3.4.

The effect of chemical

The time taken to remove the type A deposit was determined from each series of images, and is displayed in Fig. 9. For each

372

food and bioproducts processing 8 8 ( 2 0 1 0 ) 365374

Fig. 7 Removal proles of U and area for type A deposit at 0.5 m s1 . Rinsing at 20 C using (a) water, and (b) Advantis 210; and rinsing at 70 C using (c) water, and (d) Advantis 210.

Fig. 8 Area proles for rinsing type A deposit with water at 0.26, 0.4, and 0.5 m s1 for temperatures of (a) 20 C, (b) 30 C, (c) 50 C and (d) 70 C. Data shown are average of three repeats, with standard deviation shown as error bar.

food and bioproducts processing 8 8 ( 2 0 1 0 ) 365374

373

Fig. 9 The cleaning times for type A deposit determined by eye using 2 wt% Advantis 210 at temperatures of 20, 30, 50, and 70 C and ow velocities of 0.26, 0.4, and 0.5 m s1 . data point there were at least two repeats. Results are plotted as averages of these repeats and error bars of standard deviation added. Chemical cleaning using 2 wt% Advantis 210 at 20 and 30 C decreased the cleaning time. The cleaning time decreased as the ow was increased from 0.26 to 0.4 m s1 but then showed no signicant change as the ow was increased further to 0.5 m s1 . This would suggest that for this deposit a velocity exists where chemical action is most effective, and beyond which little increase in removal is not seen. Increasing the temperature from 20 to 30 C decreased the cleaning time at all ow velocities. Chemical cleaning using 2 wt% Advantis 210 at 50 and 70 C gave a further decrease in the cleaning time with increasing ow velocity. Chemical rinsing at 70 C revealed the shortest cleaning times for all ow velocities. Similar cleaning times were observed when rinsing at 30 and 50 C suggesting that raising the temperature from 30 to 50 C would not decrease cleaning time, and that suggests there is a range of temperatures where chemical action may not be enhanced by increasing the temperature. In industrial practice, if a lower temperature can give similar cleaning efciency this would be preferred as it would lead to a cost reduction and lower environmental impact. Industrial practice is often to use cleaning uid at 70 C, but this work suggests that lower temperatures may be acceptable if time is not critical.

industry currently use. i.e. that there is a cope for cleaning regimes that would have a lower environmental impact and cost. Comparable cleaning times for rinsing at 30 and 50 C also indicated there is a range of temperatures where chemical action cannot be further enhanced by increasing the temperature. This suggests cleaning at lower temperatures, i.e. 3050 C, if a CIP regime is not time limited. Chemical cleaning at 70 C showed the shortest cleaning times for all ow velocities. To reduce the cleaning requirement of the detergent phase of cleaning the data suggests the most efcient temperature is 50 C. The detergent phase could then be carried out at a lower temperature and/or concentration to remove the foulants. The next stage of the work will be to test the applicability of these ndings in an industrial fermenter and to investigate the effect of decreasing the chemical concentration.

Acknowledgements
The authors wish to thank the EPSRC and Heineken UK Ltd for provision of an EngD studentship to KRG. This paper reports results from the ZEAL project TP//ZEE/6/1/21191, which involves; Alfa Laval, Cadbury Ltd., Ecolab Ltd., Newcastle University, Scottish & Newcastle Ltd., GEA Process Engineering Ltd., Unilever UK Central Resources Ltd., Imperial College of Science Technology and Medicine, GlaxoSmithKline, Bruker Optics Ltd. and the University of Birmingham. The project is co-funded by the Technology Strategy Boards Collaborative Research and Development programme, following open competition. For more information visit http://www.innovateuk.org.

References
Akhtar, N., Asteriadou, K., Robbins, P.T., Zhang, Z., Fryer, P.J., 2010. Matching the nano- to the meso-scale: experiments with atomic force microscopy and micromanipulation. Trans IChemE C. Aziz, N.S., 2008. Factors that affect cleaning process efciency, School of Chemical Engineering, PhD Thesis, University of Birmingham. Benson, R.S., McCabe, D.J., 2004. From good manufacturing practice to good manufacturing performance. Pharmaceutical Engineering 24, 2634. Chandra, J., Huhn, D.M., Mukherjee, P.K., Hoyer, L.L., McCormick, T., Ghannoum, M.A., 2001. Biolm formation by the fungal pathogen candida albicans: development, architecture, and drug resistance. Journal of Bacteriology 183, 53855394. Christian, G.K., 2004. Cleaning of carbohydrate and dairy protein deposits, School of Chemical Engineering, PhD Thesis, University of Birmingham. Christian, G.K., Fryer, P.J., Liu, W., 2006. How hygiene happens: physics and chemistry of cleaning. International Journal of Dairy Technology 59, 7684. Cluett, J.D., 2001. Cleanability of certain stainless steel surface nishes in the brewing process, School of Mechanical Engineering, PhD Thesis, Rand Afrikaans University. Cole, P.A., Robbins, P.T., Owen, E.G., Asteriadou, K., Fryer, P.J., 2010. Comparison of cleaning of toothpaste from surfaces and pilot scale pipework. Trans. IChemE C. EHEDG Yearbook, 2009. Trends in Food Science and Technology 20 (Suppl. 1), S1S110. Friis, A., Jensen, B.B.B., 2002. Prediction of hygiene in food processing equipment using ow modelling. Trans IChemE C 80, 281285. Fryer, P.J., Asteriadou, K., 2009. A Prototype cleaning map: a classication of industrial cleaning processes. Trends in Food Science & Technology 20, 225262.

4.

Conclusions

Rheological characterisation of yeast foam has revealed that yeast slurry can be used to mimic industrial yeast deposit. The deposit viscosity and yield stress might be increased to be more industrially representative by increasing the aging time of the deposit. Studying the viscoelastic properties of the deposits also revealed that the deposit became more elastic when heated to 70 C. This deposit also takes longer to be removed by the ow. The use of rheology in predicting cleaning behaviour should be further investigated. Measuring deposit reduction by estimating the area from images has proved effective for these experiments, however an indication of volume removal would be more useful in capturing deposit swelling and erosion phenomena. Chemical rinsing using 2 wt% Advantis revealed a visually clean surface at all temperatures and ow velocities investigated. This nding suggests that a visually clean surface can be achieved at lower temperatures and ow velocities than

374

food and bioproducts processing 8 8 ( 2 0 1 0 ) 365374

Gillham, C.R., Fryer, P.J., Hasting, A.P.M., Wilson, D.I., 2000. Enhanced cleaning of whey protein soils using pulsed ows. Journal of Food Engineering 46, 199209. Gell, C., Czekaj, P., Davis, R.H., 1999. Microltration of protein mixtures and the effects of yeast on membrane fouling. Journal of Membrane Science 155, 113122. 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 and Surfaces Part B: Biointerfaces 49, 126135. Henningsson, M., Regner, M., stergren, K., Trgrdh, T., Dejmek, P., 2007. CFD simulation and ERT visualization of the displacement of yoghurt by water on industrial scale. Journal of Food Engineering 80, 166175. Liu, W., Aziz, N.A., Zhang, Z., Fryer, P.J., 2007. Quantication of the cleaning of egg albumin deposits using micromanipulation and direct observation techniques. Journal of Food Engineering 78, 217224. Mercier-Bonin, M., Ouazzani, K., Schmitz, P., Lorthois, S., 2004. Study of bioadhesion on a at plate with a yeast/glass model system. Journal of Colloid and Interface Science 271, 342350. Mores, W.D., Davis, R.H., 2002. Yeast foulant removal by backpulses in crossow microltration. Journal of Membrane Science 208, 389404.

Mozes, M., Marchal, F., Hermesse, M.P., van Haecht, J.L., Reuliaux, L., Leonard, A.J., Rouxhet, P.G., 1987. Immobilization of microorganisms by adhesion: Interplay of electrostatic and nonelectrostatic interactions. Biotechnology and Bioengineering 30, 439. ORourke, T., 2003. CIP Cleaning In Place. The Brewer International 3, 3034. Othman, A.M., Asteriadou, K., Robbins, P.T., Fryer, P.J., 2010. Cleaning of sweet condensed milk deposits on a stainless steel surface. In: Wilson, D.I., Chew, Y.M.J. (Eds.), Proceedings of the Fouling & Cleaning in Food Processing. 2010 Conference in Cambridge, Session V. Department of Chemical Engineering, Cambridge, pp. 174182. Sahu, K.C., Valluri, P., Spelt, P.D.M., Matar, O.K., 2007. Linear instability of pressure-driven channel ow of a Newtonian and a HerschelBulkley uid. Physics of Fluids 19, 122101. Salo, S., Friis, A., Wirtanen, G., 2008. Cleaning validation of fermentation tanks. Food and Bioproducts Processing 86, 204210. Tamime, A.V., 2008. Cleaning-in-place: Dairy, Food and Beverage Operations, Society of Dairy Technology Series. Wiley-Blackwell, London.

Potrebbero piacerti anche