You are on page 1of 37

JOURNAL OF AEROSOL MEDICINE Volume 16, Number 4, 2003 Mary Ann Liebert, Inc. Pp.

341377

Invited Paper Cascade Impactors for the Size Characterization of Aerosols from Medical Inhalers: Their Uses and Limitations
JOLYON P. MITCHELL, Ph.D., and MARK W. NAGEL, H.B.Sc.

ABSTRACT Cascade impactors, including the multi-stage liquid impinger, are by far the most widely encountered means for the in vitro determination of the particle size distribution of aerosols from medical inhalers, both in product development, batch release and in applications with add-on devices. This is because they directly measure aerodynamic size, which is the most relevant parameter to describe particle transport within the respiratory tract. At the same time, it is possible to quantify the mass of active pharmaceutical ingredient in different size ranges independent of other non-physiologically active components of the formulation. We begin by providing an overview of the operating principles of impactors and then highlight the various configurations and adaptations that have been adopted to characterize the various classes of inhaler. We continue by examining the limitations of the cascade impaction method, in particular looking at potential sources of measurement bias and discussing both appropriate and inappropriate uses of impactor-generated data. We also present a synopsis of current developments, including the Next Generation Pharmaceutical Impactor, and automation of cascade impactors for routine inhaler performance measurements. Key words: impactor, impinger, inertial fractionator, medical inhaler, particle size analyzer

INTRODUCTION

(CIs), including the multistage liquid impinger (MSLI), are the most commonly encountered group of instruments for in vitro size-analyzing aerosols produced by medical inhalers. They are the equipment of choice in both U.S. and European Pharmacopeias, 1,2 and are also recommended in current guidance documents for industry published by the correASCADE IMPACTORS

sponding regulatory authorities. 3,4 The exception is the size-characterization of aqueous nasal spray-pumps, since these inhalers generate droplets that are in general larger than the measurement range capability of CIs. At a fundamental level, a CI should not be considered an in vitro lung simulator, as it operates at constant flow rate, in contrast with the continually varying flow rate associated with the breathing cycle. Nevertheless, this class of in-

Trudell Medical International, London, Ontario, Canada.

341

342

MITCHELL AND NAGEL

strument, if used correctly, can provide particle size information, based on aerodynamic diameter (dae), which may be indicative of the likely deposition of active pharmaceutical ingredient (API) in the respiratory tract. dae takes into account the influences of both particle density and shape and is related to particle physical size (dp ) measured by microscopy through the equation: rp C p C ae dae 5 dp } xr0

nasal cavity.11 Elutriators that operate on the principle of sedimentation have been developed for size separating larger particles, 12 but these devices have not so far been widely applied to the testing of either nasal or oral inhalers.

OPERATING PRINCIPLES
A typical CI comprises several stages, each of which functions as a size-separator or fractionator of the incoming aerosol in a gas stream moving at constant velocity (flow rate). In concept, a single stage impactor comprises a jet or nozzle plate containing one or more circular or slotshaped orifices located a fixed distance from a collection surface that is usually horizontal (Fig. 1). The stage functions by classifying incoming particles of various sizes on the basis of their differing inertia, the magnitude of which reflects the resistance to a change in direction of the laminar flow streamlines.5 As the incoming flow passes through the nozzle plate, the streamlines diverge on approach to the collection surface, whereas the finite inertia of the particles causes them to cross the streamlines. The dimensionless Stokes number (St), which is the ratio of the stopping distance of a particle to a characteristic dimension, in this case the nozzle diameter, W (or average diameter, for a multi-orifice stage), describes the process, defining a critical particle size that will reach the collection surface for a particular stage geometry.5 The theory underlying impactor function has been developed over the past 25 years by solving the Navier-Stokes equations defining the gas flow field in the absence of particles, and then using Newtons equation of motion to model the passage of different sized particles through various stage geometries. 1315 Marple et al.9 have recently summarized the current status of impactor theory, so that only the essentials are given

1/2

(1)

where the Cunningham slip correction factors (Cae and Cp ) are close to unity when dp . 1.0 mm, and the dynamic shape factor (x) is unity for spherical particles. 5 If the model of Rudolph et al., 6 which assumes oral tidal breathing, is chosen as an example, depending on inhalation flow rate, particles with dae larger than approximately 68 mm deposit mainly in the oropharyngeal region, central (bronchial) airway deposition peaks with particles having dae of 79 mm, and peripheral (alveolar) deposition in the lung reaches a maximum with particles having dae of 24 mm.7 Since CIs that operate at ambient pressure are efficient size-separators in the range of 0.312 mm in aerodynamic diameter, 5 they are well suited for characterizing almost all oral inhaler-produced aerosols. Although particles with dae finer than 0.3 mm can be analyzed with adequate resolution by CIs having some stages operated at sub-atmospheric pressure (so-called low-pressure impactors),8 such ultrafine particles are of little interest in the context of inhaler characterization, because they contain almost no mass of API, and furthermore they are likely to be exhaled if they do not deposit either by phoretic or diffusion-based mechanisms.7 Such CIs will therefore not be discussed further. Of greater significance from the standpoint of inhaler testing is the influence of gravitational sedimentation on the motion of larger particles, which imposes the upper size limit for CI measurements in the region of 2025 mm in aerodynamic diameter. 9 The increasing effect of gravity on the inertial size-separation of larger particles has so far proved to be a barrier to impactor development, and in consequence, commercially available CIs have little size-discriminating capability for particles larger than this limit. 9 They are therefore unsuitable to characterize aqueous droplet sprays produced by mechanical pump-driven nasal drug delivery devices,10 other than to quantify the mass fraction, typically defined as being ,10 mm aerodynamic diameter, that is likely to penetrate beyond the

FIG. 1. Cross-section through a single stage impactor illustrating principle of operation.

CASCADE IMPACTORS AND SIZE OF AEROSOLS FROM INHALERS

343

here. These concepts have been applied to the design of the Next Generation Pharmaceutical Impactor (NGI, MSP Corp., Shoreview, MN).16 For a single nozzle (jet) impactor, St is related to W through the expression:
2 r p Cp d 2 U p r0C aed aeU St 5 } 5 } } 9 mW 9 mW

(2)

in a self-consistent set of units, based on either particle physical or aerodynamic diameter. Equation 2 predicts that the particle collection efficiency (E) of an ideal impactor stage, expressed as a percentage, will increase in a stepwise manner between limits of zero to 100%. In practice, for a well-designed stage, E is a monotonic sigmoidal function of St or dae that increases steeply from E of <0% to .95%, reaching its maximum steepness when E is 50% (Fig. 2). At this location, defined as the cut size (d50): 9hW C 50 d50 5 } r0 CaeU 9pmnW3 C50 d50 5 } 4r0CaeQ

1/2

St50

for a multi-orifice stage comprising n circular nozzles. It is possible to take into account the shape of the actual collection efficiency curve of the stage in the analysis of impactor data,17 but this refinement is rarely done for measurements of inhaler performance. Instead, the assumption is made that the mass of particles larger than d50 (the size corresponding to E50), that penetrate the stage, is exactly compensated by the mass associated with particles finer than this size, that are collected. Thus, the cut size can be defined as a single valued constant for a given stage at a fixed flow rate. Particles with of dae of $ d50 are assumed to be fully collected, whereas all particles finer than d50 are deemed to penetrate the stage. The so-called sharpness-of-cut of the stage can be defined as the geometric standard deviation (GSDstage ) of the efficiency curve by analogy with the properties of the log-normal distribution function: GSDstage 5

(3)

or in terms of volumetric flow rate (Q):

1/2

d84.1 } d15.9

(5)

St50

(4)

GSD stage for a well-defined stage is ideally less than 1.218 (a GSDstage of unity corresponds to the

FIG. 2. Impactor collection efficiency curve, showing parameters characterizing stage performance.

344

MITCHELL AND NAGEL

ideal size-separator). However, with many existing designs of CI, GSD stage values are in excess of this limit, particularly for those stages that sizefractionate particles larger than about 5 mm in aerodynamic diameter where gravitational settling contributes significantly to the size-separation process.19 Marple and Liu13 and Rader and Marple, 15 have identified that the value of St at E50 (St50 ) should be close to 0.49 for well-designed round-nozzle impactors, where differences in particle inertia dominate the size separation process. However, at least two other parameters also appear to be important for effective size-separation. The ratio of nozzle-to-collection surface distance/nozzle diameter (S/W) describes the geometry of the stage. St50 (and hence d50 ) is unaffected by small variations in S, if S/W . 1.0. At the same time, the dimensionless flow Reynolds number (Ref), defined as: ra UW Ref 5 } m (6)

erwise be collected are instead captured by the cross-flow and transferred beyond the collection surface. Increased inter-stage losses as well as bias towards smaller sizes can therefore result with multi-stage impactors.20 The dimensionless cross-flow parameter (Xc), is defined as: nW Xc 5 } 4Dc (7)

should be in the range of 5003000 to optimize GSD stage . Further unpublished investigations have indicated that S/W can be between 1 and 10 for effective size-separation to occur, allowing considerable latitude in stage design.20 There is therefore some freedom in establishing the precise location of the collection surface beneath the nozzle-plate with many impactor designs. More importantly from a users standpoint, the presence of a coating to improve particle collection behavior is unlikely to affect stage performance. The ratio of nozzle throat length (T) to W can also influence size separation efficiency, 13 decreasing St50 with increasing T/W. However, the effect is likely to be small with commercially available impactors, where T/W is typically ,10.20 Even though these criteria provide the basis for a sound impactor design, it is notable that S/W is ,1.0 for both stages 0 and 1, and Ref is below 500 (110 , Ref , 394) with stages 0 to 6 of the widely used Andersen 8-stage CI ([ACI] ThermoAndersen Inc., Smyrna, GA) at the design flow rate of 28.3 L/min (1 ACFM).21 Cross-flow induced by air exiting the nozzles near the center of the nozzle plate and flowing outward past other air jets located near the periphery of the nozzle cluster, can prevent the air jets near the edge of the cluster from reaching the impaction plate with multi-nozzle designs.22 Under these circumstances, particles that would oth-

X c should be ,1.2 to avoid cross-flow related problems, and this criterion is achieved for all of the commonly encountered CIs used with inhaler testing, with the exception of stage 2 of the ACI, where Xc is 1.2. 21 Several individual impaction stages are connected together in a CI, most commonly in a vertical stack, but it is notable that the NGI has all of its stages located adjacent to each other in the horizontal plane, primarily for ease of use in semi- or fully-automated operation.16,18 Since the purpose of a CI is to fractionate the incoming aerosol into progressively finer particle sizes, beginning with the coarsest particles, U is increased from one stage to the next, primarily by reducing the nozzle diameter. The number of nozzles per stage as well as the number of stages within the CI can also be adjusted to optimize size resolution. However, in practice, inserting more than five stages per decade of particle size is counterproductive, because the stages immediately before a given stage will interfere with efficient particle collection, due to the non-ideal nature of their collection efficiency curves. This restriction limited the number of stages of the NGI with cut sizes in the range of 0.55 mm in aerodynamic diameter to five.16

DATA INTERPRETATION FROM CIs


Fundamentally, CIs determine particle aerodynamic mass-weighted size distributions 5,9 from which several parameters can be derived in order to quantify inhaler performance. The MMAD, as the measure of central tendency of the size distribution, is the most often reported value. Berg et al.23 have suggested that, for the purpose of standardizing inhaler testing, the MMAD should be based on that portion of the dose entering the CI, rather than the total emitted dose ex inhaler actuator. This approach is valid if the intention is to use the size distribution data to estimate lung deposition. However, the determination of MMAD should include the components of the

CASCADE IMPACTORS AND SIZE OF AEROSOLS FROM INHALERS

345

dose depositing in the induction port (and preseparator if used), when the purpose of the measurements is to evaluate add-on devices for inhalers that are intended to remove the portion of the dose likely to deposit in the oropharynx.24 In addition to determining MMAD, it is often helpful to have a single parameter that quantifies the spread of the size distribution. If the portion of the dose entering the impactor is considered, the size distribution can often be approximated to a log-normal function (particularly with solution pressurized metered-dose inhaler (pMDI)based formulations,25,26 and the spread of the distribution, defined by the geometric standard deviation (GSD), can be estimated as: GSD 5

d84.1 } d15.9

(8)

where d15.9 and d84.1 are the sizes corresponding to the mass-percentile values of 15.9% and 84.1%, respectively, for the cumulative size distribution. A degree of caution has to be exercised in the use of GSD values, since there is no fundamental reason why inhalers should generate aerosols having log-normal particle size distributions. For instance, the coarse ballistic fraction results in a second mode in the size distribution of pMDI-generated aerosols when the entire dose from is considered.23 Deviations from log-normality have also been observed with size distribution data based only on mass entering the CI for at least one pMDI-generated formulation, in this instance being associated with a large number of drug free surfactant particles. 27 Bi-modality has been reported when considering the entire dose leaving the inhaler with certain carrier-based DPI-generated aerosols,28 rendering the use of GSD as a descriptor inappropriate unless some restriction is placed on the size range being considered.23 In an effort to minimize the amount of data required for the description of inhaler-based aerosols, Thiel,29 like Berg et al., 23 proposed separating the dose captured in the induction port and pre-separator (if used), from the non-ballistic fraction (NBF) entering the impactor. MMAD and GSD were both used to define the portion of the dose entering the CI, as has already been described. The NBF became the fraction of the total emitted dose ex inhaler mouthpiece that comprises the area under the log-normal curve defined by the MMAD and GSD. Thiel29 then used values of MMAD, GSD and NBF, together with the inhaled volume, inhalation rate and breathhold, and entered into the Stahlhofen-Rudolph

model of lung deposition,6 to show that these parameters have predictive value for lung deposition for representative examples of both pMDIs and DPIs, supported by g-scintigraphic evidence from four separate clinical studies. In the context of judging similarity of CI-measured particle size distributions from different inhalers, a working group at the Product Quality Research Institute (PQRI) has begun exploring the use of statistical methods based on modified x2 analyses of CI-stage data to develop a robust metric for their comparisons.30 PQRI is a collaborative process involving representatives from the pharmaceutical industry, academia and the FDA, and the focus of this activity is concerned with comparing inhalers from manufacturers of innovator and generic formulations for regulatory purposes. Their methodology is only applicable where large databases from CI measurements are available, but it may offer a means to compare test and reference products with a metric that does not depend on particular product type or particle sizing equipment. However, Clark and Kadrichu31 also investigated the applicability of both the f2 similarity factor and x2 statistic in a comparison of ACI- and MSLI-generated size distribution data with several reference distributions having MMAD values in the range of 18 mm with a constant GSD of 1.2. By subsequently computing lung deposition data for the reference and test distributions using the Stahlhofen-Rudolf model,6 they showed that neither statistic responds to changes in size distribution in a meaningful way in terms of lung deposition. As an alternative methodology, they developed the concept of theoretical deposition fraction (TDF) of the dose, obtained by multiplying the mass-percentage of API collected on each CI stage by the lung deposition probability based on the mean diameter for that stage. They validated this approach by demonstrating a sensitive correlation between TDF and differences in stage deposition patterns in ACI and MSLI data between laboratories that had participated in a round-robin inter-laboratory study using a commercially available albuterol pMDI inhaler. 32 In the light of these findings, further work is needed to establish the applicability of appropriate statistical methodologies for comparing CI-generated size distribution data from other inhalers. For many purposes, it is unnecessary to go to such lengths to compare data from CI-testing, and it is sufficient to determine a sub-fraction of the dose entering the CI that is appropriate for

346

MITCHELL AND NAGEL

the therapeutic class of the API (fine particle fraction [FPF]). The European Pharmacopeia specifies the upper size limit of 5 mm in aerodynamic diameter for FPF,2 which very often necessitates interpolation of the cumulative size distribution data to arrive at the required value. In contrast, the U.S. Pharmacopeia does not specify precise size limits for FPF, allowing the tester to chose range limits appropriate for the formulation/inhaler being tested.1 This approach recognizes that formulations in different therapeutic classes may have optimum efficacy in different size ranges. In practice, the upper size limit for FPF is set at either 5.8 or 4.7 mm, corresponding to two cut-point sizes of the widely used ACI at 28.3 L/min to avoid interpolation of the size distribution data. The compendial procedures only define the upper size limit, and therefore an underlying assumption is made that the entire content of the dose collected in the CI smaller than this size comprises fine particles likely to deposit in the lungs. This supposition may be appropriate for products where almost all of the dose emitted from the inhaler is contained in particles in the size range from 1 to 5 mm in aerodynamic diameter. However, there is some concern about the treatment of some newer solution-based pMDIdelivered formulations, in which as much as 40% of the mass that enters the CI can be contained in particles with dae , 1 mm. A significant portion of such extra-fine particles may be exhaled without depositing in the lung.7 In response to this concern, the reporting of performance data based on two sub-fractions (fine component having dae , 4.7 mm [FPF4.7 mm ] and extra-fine component with dae , 1.1 mm [EPF1.1 mm ]) has been proposed in a recently developed Canadian Standard for spacers and holding chambers.24 The literature contains numerous other examples of similar sub-fraction definitions, for instance those of Dolovich26 (three sub-fractions corresponding to particles with dae , 1, 3, and 5 mm). However, it should be noted that FPF defined in this way still includes the entire CI contents beyond the appropriate stage defining the upper limit, rather than adopting a more logical approach based on likelihood of lung deposition, by defining FPF in terms of a range between appropriately defined size limits. Some formulators therefore group CI stages to represent that portion of the dose most likely to reach the target receptors, for example considering the mass collected on stages 35 of an ACI operated at 28.3 L/min, and thereby rep-

resenting particles 1.14.7 mm in aerodynamic diameter. 33 The use of the term lung targetable fraction (LTF) suggested by Newhouse34 in the context of interpreting in vitro data is interesting, since this definition is based in part on the outcome from a small clinical investigation, rather than from attempts at predicting lung deposition based on models. This study involved the delivery of a pMDI-based bronchodilator via a holding chamber to 20 patients with mild-to-moderate and moderate-to-severe asthma. Newhouse34 found that particles with dae , 2 mm were most relevant to predict the likely response in terms of the widely used metrics forced expiratory volume in 1 sec and maximum expiratory flow of 2575% of vital capacity, following deposition at the receptors in the lungs. A limit for LTF at finer particle sizes was not reported, and the applicability of this definition to other therapeutic classes of inhaled medication and patient categories (infant/child/adult) is unknown. However, this study established the principle of establishing a link between sub-fraction limits and physiological response, which is a desirable goal to attain where it is possible to do so. More than particle size information can be extracted from CI-based measurements. However, there are concerns about the value of such metrics to quantify measurement quality as well as inhaler performance. Perhaps the most contentious issue is the CI-derived material (mass) balance. In the compendial methods,1,2 the API material balance from the inhaler actuator/ mouthpiece, induction port and CI is required to validate that the equipment operated correctly and that the inhaler functioned when actuated. This additional piece of data can therefore provide important information about the quality of the measurement, but it is relatively imprecise compared with the direct measurement of dose (total emitted dose ex actuator) by the procedures for determining inhaler content uniformity. A CIgenerated material balance is composed of several individual measurements in which sources of variability (e.g., sample preparation, analytical precision) accumulate, compared with the single measurement of API mass collecting on the filter and interior surfaces of the dosage unit sampling apparatus. Current FDA Guidances to Industry for inhaler product quality control3,35 indicate that the mass of API recovered in the CI and accessories should be within 615% of the label

CASCADE IMPACTORS AND SIZE OF AEROSOLS FROM INHALERS

347

claim dose on a per actuation/spray basis. These limits are comparable with those recommended for dose content uniformity, based on the mean for each of the beginning and end determinations. Poochikian and Bertha36 have expressed the view that, whereas the material balance by itself does not provide information on the size distribution of the emitted dose from the inhaler, in conjunction with dose content uniformity testing, it does give additional reassurance that both the CI functioned properly and that the analytical methodology was performed correctly. Furthermore, if the CI-generated material balance is inconsistent with the equivalent data from dose content uniformity testing, this outcome should signal a potential problem with the CI system. Although it is generally accepted that the CI-material balance cannot be thought of as a system suitability criterion, because of the absence of reference standards against which to establish CI accuracy and reproducibility, 36 there are widespread concerns within the pharmaceutical industry that the proposed limits are too restrictive, supported by the outcome of a retrospective analysis of an extensive database including both pMDI- and DPIbased formulations.37 The relative imprecision of the CI-material balance, together with the perception of significant producer-risk that a failure to achieve the specified range might result in a batch failure, when the cause was a fault with the CI methodology, has resulted in a comprehensive reappraisal of the causes of CI-material balance failures and the appropriateness of this metric as a batch release specification. 38 This work is being undertaken at PQRI, where a working group has developed an analysis of the various causes of failures for both material balance and particle size distribution measurements by the CI method.39 This consensus about what constitutes so-called Good Cascade Impactor Practice is part of a concerted effort to develop the most appropriate use of the CI-generated material balance in the context of inhaler product testing. In this way, diagnosis of the cause of a material balance failure should be able to be undertaken in a logical manner. Since total emitted dose from the inhaler (TED) is also obtainable from a CI determination, this parameter is often used in the presentation of in vitro performance data. The additional information is particularly useful for the assessment of add-on devices with pMDIs,40,41 where it may be important to have an indication of the magnitude

of reduction of the coarse component brought about by the spacer or holding chamber. Outside of the context of inhaler quality control testing, it can be argued that TED may better be measured with the inhaler attached to a breathing simulator, than by sampling at constant flow rate into a CI. This supposition is certainly true when testing holding chambers at low flow rates, where the behavior of inhalation and exhalation valves can only be properly evaluated by simulating the full respiratory cycle.42 Whether CI or breathing simulator alone is used to establish TED, the fine particle dose (FPD) based on an appropriate upper size limit can be calculated from CI-based data in accordance with: FPD 5 TED 3 FPF (9)

However, it should be noted that if TED is obtained from CI measurements, then both FPD and TED are also subject to loss of precision from the same cause as has been described for the material balance. Grouping the mass of API collected on several stages to represent a therapeutically significant size range33 does not eliminate this loss of precision, unless the recovered API from the stages concerned is combined at assay. Under these circumstances, variability may be reduced, arising from individual stages on which the mass of API collected is close to the limit of detection by the assay procedure.

CIs IN USE WITH INHALER TESTING


Although there are many different types of CI in use for aerosol characterization, the vast majority of inhaler testing is undertaken with the impactors that are listed in the compendia (Table 1), and the reader is referred to these publications for detailed information about their design. The stage characteristics including values of Ref, S/W and d50 for the ACI at 28.3 L/min, 150/160-series Marple Miller impactors (MMI) at 30 and 60 L/min, and the MSLI at 60 L/min, corresponding to flow rate-CI combinations widely used for inhaler characterization, have been summarized by Marple et al.21 Well-designed comparisons of different CIs in current use with a range of inhaler types are few, and in this context, the study by Olsson et al.43 is significant. They compared size distribution data for four different formulations (3 DPIs and 1 pMDI) measured by ACI, 160

348
TABLE 1. COMPENDIA L CIS Impactor Andersen 8-stage (ACI)no pre-separator Marple-Miller series 160 (MMI) Andersen 8-stage (ACI)pre-separator Multi-Stage Liquid Impinger (MSLI) Next Generation Pharmaceutical Impactor (NCI)
a Proposals

MITCHELL AND NAGEL

US Pharmacopeiaa Apparatus 1 Apparatus 2 Apparatus 3 Apparatus 4 Apparatus 5 Apparatus 6 for pMDIs for DPIs for DPIs for DPIs for DPIs for pMDIs

European Pharmacopeiab Apparatus D Apparatus D Apparatus C Apparatus E

for the NCI. A and B of the European Pharmacopeia are the two-stage glass and metal twin impingers, respectively. The status of these devices is currently under review.
b Apparatuses

series MMI and five-stage MSLI, each operated at 60 L/min, observing that the MMI and MSLI gave well correlated and similar size distribution data for all the inhalers. However, the outcome of an examination of an extensive database of ACI- and MSLI-measured data from the budesonide Turbuhaler DPI (n 5 40 measurements) reported in the same paper, indicated that ACI-based size distributions were consistently narrower (less polydisperse (GSD < 1.4)) than those obtained with the MSLI (GSD < 2.0). Nevertheless, MMADs measured by both CIs were quite similar, being in the range of 2.32.8 mm (the ACI-based values clustering towards the upper end of the range with the MSLI values at the lower end). Subba Rao et al.44 also observed a similar finding in a comparison of size distributions following 10 actuations from a pMDI-delivered peptide suspension formulation measured by ACI and 150 series MMI at 28.3 and 30 L/min, respectively. In their case, the GSDs were 1.8 and 1.95 using the ACI and MMI, respectively; however, the MMAD determined by MMI (4.0 mm) was significantly larger than the corresponding ACI-measured MMAD (2.9 mm). They attributed these differences to bias caused by higher inter-stage (wall) losses of larger particles in the ACI, which resulted in an over-estimation of the proportion of finer particles with this CI. Larger losses were also observed with the ACI by Olsson et al., 43 but not quantified in terms of particle size. Although Subba Rao et al.44 concluded that, for batch control purposes, absolute equivalency of these CIs was not critical, such a conclusion would seem to be less apt, if the intention behind the particle size measurement is ultimately to estimate lung deposition. Perhaps the advice of Olsson et al., 43 that caution should be exercised when comparing results from different CIs, is the most appropriate stance to take, given

the variety of different purposes underlying these measurements. Besides the standard CI configurations, there are some important variants that are used to encompass the entire range of flow rates likely to be encountered with inhaler testing. It should be noted that the ACI was developed as a room-air bacteriological sampler, operating at a fixed flow rate of 28.3 L/min.45 This impactor in its standard form is therefore unsuitable for use at flow rates significantly in excess of this value, as is required for the testing of many DPIs. Recognizing this limitation, Nichols et al.46 undertook modifications that enable the ACI to be used at 60 L/min by the removal of stage 7 and insertion of a new higher flow rate stage 21. Stage 21 has larger nozzles than stage 0 to permit the cut size to be maintained close to 9.0 mm in aerodynamic diameter at the higher flow rate. Stage 7 is removed, as its cut size at 60 L/min would be too fine to be of much use with inhalers. The stage calibration data presented with the modified ACI indicated comparable size selectivity to the standard ACI for stages 04 (GSD values 1.251.4).46 However, although the sharpness of cut for stage 21 is comparable with that for stage 0, there appears to be considerable overlap of the two collection efficiency curves, suggesting that the new stage may be interfering with the collection of particles that ought to reach stage 0. More recently, a further similar development of the ACI has taken place, replacing stage 6 by stage 22. This modification has enabled the impactor to be used at flow rates up to 90 L/min,47 which is particularly applicable for testing low resistance DPIs. However, the pre-separator cannot be used with either of these modifications to the ACI, since its cut size, which is close to 9.0 mm in aerodynamic diameter at 28.3 L/min,48 is significantly finer than the cut size of the first impaction stage at the

CASCADE IMPACTORS AND SIZE OF AEROSOLS FROM INHALERS


TABLE 2. STAGE CUT SIZES (mM ) FOR ACI AT SELECTED FLOW RATES Flow rate (L/min)a Stage 22 21 0 1 2 3 4 5 6 7 28.3 Not used Not used 9.0 5.8 4.7 3.3 2.1 1.1 0.7 0.4 60 Not used 8.60 6.50 4.40 3.20 1.90 1.20 0.55 0.26 Not used 90 8.00b 6.50b 5.20b 3.50b 2.60b 1.70b 1.00b 0.22b Not used Not used
THE

349

a Data at 60 and 90 L/min from ref. 47 (nominal values from the manufacturer for 28.3 L/min). bCut size reported by Thermo Andersen for this stage is 0.43 mm.

higher flow rates. Stage cut sizes at 60 and 90 L/min based on measured values for both modified versions of the ACI47 are summarized in Table 2, together with the nominal values supplied by the manufacturer for the standard configuration operated at 28.3 L/min. All configurations of the ACI make use of circular-shaped flat metal or glass discs (plates) to collect particles. A filter circle supported on the (inverted) plate of each stage can also be used as a collection substrate, but this option is seldom used for inhaler testing on account of increased complexity for API recovery. There is also evidence that filter substrates modify significantly the stage collection characteristics of the ACI as a result of flow penetrating the porous filter matrix.49,50 The simple geometry associated with a rigid metal or glass plate makes recovery of the API a comparatively easy process compared with collection surfaces that have depth. The MMI series of five-stage impactors are currently available in three sizes, all based on the original work of Miller 51 and Marple et al., 52 enabling measurements to be made at flow rates of 4.990 L/min with cut sizes that are all located within the useful range for inhaler testing (Table 3). Stage collection efficiency curves for all versions of this impactor expressed either in terms of dae or St are steep, and associated with GSD stage values that are close to or below 1.2.5153 These impactors employ collection cups as particle collectors, with the purpose of improving productivity. However, API recovery can be more difficult than with the simpler geometry of col-

lection plates with recovery procedures requiring more than contact with solvent to dissolve the collected particles. The model 160 (high flow) MMI is the standard configuration intended for use at 6090 L/min. The model 150 MMI has half the number of nozzles per stage compared with the model 160 MMI, for use as an alternative to the ACI for pMDI-characterization at 3060 L/min. A low flow rate version (model 150P) was developed to permit pMDIs with add-on devices intended for low flow patients to be tested at more appropriate conditions of use (4.9 and 12 L/min).53 Although these impactors have not been widely adopted, internal losses for model 150 and 160 MMIs reported by Marple et al., 52 were no more than 5% of the incoming aerosol at worst case (4 , dae , 6 mm), decreasing to ,1% for finer particle sizes and ,2% for larger particles. These measurements were based on calibration with monodisperse droplets. The model 160 MMI has subsequently been reported as having internal losses at 60 L/min with at least one DPI (Bricanyl Turbuhaler) 54 that were comparable with those indicated by Marple et al., 52 provided that precautions were taken to eliminate particle bounce and re-entrainment by coating the collection surfaces with a tacky surface (silicone oil). Losses within the low flow MMI have also been reported as being less than 5% of the material balance from two types of pMDI-generated formulations.53 The relative lack of interference between successive stages within all versions of the MMI5153 compared with that evident particularly with the upper stages of the ACI48,55 may account for the differences observed between model 160 MMI-

TABLE 3. STAGE CUT SIZES (mM ) FOR VARIOUS MODELS OF THE MMI MMI model and flow rate (L/min) Stage 1 2 3 4 5 150P 4.9 10.0 7.2 4.7 3.1 0.77 150P 12.0 10.0 4.7 3.1 2.0 0.44 150 30 10.0 5.0 2.5 1.25 0.63 160 60 10.0 5.0 2.5 1.25 0.63 160 90 8.1 4.0 2.0 1.0 0.5

Data values at flow rates other than 90 L/min are based on calibration data.46,47 Data at 90 L/min were calculated, based on the behavior of inertial collectors as a function of flow rate.

350

MITCHELL AND NAGEL

and ACI-measured size distribution data from four different albuterol pMDIs reported by LeBelle et al.56 This group observed that although GSD values were comparable between the two types of CI, MMI-measured MMADs were systematically finer, by an average of 18%. The aerodynamic characteristics of the NGI were developed from similar concepts used in the Marple-Miller series of impactors, in particular the use of collection cups rather than flat plates and nozzle configurations to minimize the impact of cross-flow.16 Calibration data obtained with a so-called archival NGI, having critical parameters (principally nozzle diameters) close to the mean values specified in the design have confirmed that the sharpness of cut for most stages is close to the design goal of 1.2 between 30 and 100 L/min.57 The micro-orifice collector (MOC) that is a substitute for the normal back-up filter may allow significant penetration of extra-fine particles with dae , 0.5 mm,16 but it can either be replaced or backed-up by a conventional filter when analyzing the few formulations containing an appreciable portion of the dose associated with such particles. The MSLI has undergone several changes during its evolution from a three-stage single nozzle device developed by May.58 The original version used for inhaler testing contained four stages including back-up filter. 59 This version was recently augmented to five stages by the addition of a lower seven-nozzle stage,60 with cut sizes determined by calibration with monodisperse particles at 60 L/min to be 13.0, 6.8, 3.1, and 1.7 mm.60 The MSLI has the advantage of not having appreciable inter-stage losses. Thus, only about 1% of DPIgenerated API was claimed as being lost to the internal surfaces of the nozzles (jet tubes),60 since the recovery solvent for the API is also used to wash down the walls of each stage. The advantage of having fewer stages compared with the ACI, from the standpoint of reducing the time per measurement, is offset by the lack of size resolution in the critical range of 0.55.0 mm in aerodynamic diameter. Furthermore the sharpness of cut (GSDstage ) for MSLI stages (in the range of 1.41.5, estimated from the reported calibration data at 60 L/min 60) is poorer than that typically achieved with either the ACI or MMI. Marple et al.21 suggested from theoretical analysis that stage D50 values for the MSLI are likely to be quite dependent on small changes in nozzle-to-collection surface distance (S), given the higher than

desirable values of Ref at 60 L/min (3.310.3 3 103 ). This prediction is borne out by the outcome of an experimental assessment of sensitivity to changes in collection plate surface level below the edge of the holder for stages 2 and 3.60 However, the presence or absence of liquid in the stages does not appear to be an important factor governing stage cut sizes. The purposes of this liquid are to wet the collection surfaces and dissolve the deposited particles. 58 Although not encountered as widely as the CIs listed in the compendia, several other impactors are occasionally encountered in the characterization of inhaler-produced aerosols, and three of the more frequently encountered instruments are worthy of further mention. The Delron or Battelle impactor, like the ACI, had its origins as a sampler associated with industrial hygiene, but has since been used to characterize inhaler-based aerosols.28,61 The DCI-6 is a six-stage vertical stack design with glass collection plates. It incorporates a back-up filter and is operated at 12.5 L/min by means of a critical orifice. The impactor is reported as having cut sizes of 11.2, 5.5, 3.3, 2.0, 0.9, and 0.5 mm.28 Stage collection efficiency data based on the calibration of a prototype impactor 62 indicate that GSD stage values are better than 1.2 for all of the stages, and therefore comparable with the performance of the various MMI designs. The 10-stage quartz crystal impactor (QCM, California Measurements Inc., Sierra Madre, CA) has been used with some success as a means for rapid screening of pMDI-based formulations.63 Although quick to use compared with conventional CIs, once the reference and measurement sensors for each stage have been properly greasecoated, its major drawback is the lack of specificity for API in the presence of solid or liquid excipients in the formulation. The operating flow rate (0.24 L/min) is very low by comparison with other CIs, making it difficult to interface this impactor with conventional induction port designs. Nevertheless, Tzou and Elvecrog64 reported comparable MMAD values for formulations that contain little or no surfactant by comparison with ACI-based data. The GSD stage values for most stages of the QCM impactor are comparable with those of the MMI, although the sharpness of cut for stage 2 is poorer than this limit, resulting in appreciable overlap with the next stage.65 The QCM has also been shown in a study with monodisperse particles carrying between 3500

CASCADE IMPACTORS AND SIZE OF AEROSOLS FROM INHALERS

351

and 4000 electronic charges to be susceptible to bias caused by electrostatic charge accumulation on the insulated stage collection surfaces, inhibiting the further collection of incoming particles.65 It may therefore be necessary to take precautions to minimize this effect when using this CI to measure the charged aerosols that are encountered with inhalers of almost all classes.6668 However, in the context of electrostatic chargerelated effects in impactors in general, it is worth noting that Dunbar and Hickey 69 did not observe significant changes to ACI-measured droplet size distributions from a jet nebulizer in experiments in which their CI stages were either electrically grounded or isolated by their inter-stage Oring seals (normal condition). This finding would suggest that accumulation of droplet-supplied electrostatic charge within the mostly aluminum ACI is unimportant. However, further work is warranted to establish if their finding is of general applicability to other types of inhaler-produced aerosols and CI types. In recent years, it has become recognized that nebulizer-produced droplets should ideally be size characterized non-invasively by light scattering techniques, such as phase-Doppler analysis69 or laser diffractometry.70,71 Currently available light scattering methods, however, do not determine the mass of API and may therefore lead to misleading results, particularly with suspension formulations where droplets may be formed without containing any drug-bearing particles. 72 Current guidance for the testing of nebulizers published by the FDA therefore require that at least one of the sizing methods for nebulizer characterization be a CI-based method.73 Dolovich70 noted that typically the volume (mass) median droplet diameter reported from laser diffractometry measurements for nebulizers is 12 mm greater than that obtained by CI analysis, attributing the cause primarily to droplet evaporation in the impactor, a process which is considered in more detail later. Dennis et al.74 proposed that aerosols should ideally be sampled at low flow rates (,15 L/min) to avoid bias due to evaporation by the entrainment of excessive amounts of drier ambient air. Even when this precaution is taken, heat transfer from conventional CIs that have high thermal mass is likely to bias droplet size measurements unless care is taken to ensure that the aerosol and CI temperatures are similar. 75,76 The Marple Model 290 Personal CI77 (Thermo Andersen Inc., Smyrna, Georgia, USA)

was therefore proposed in a new CEN standard78 as the impactor of choice for the size characterization of these aerosols, as it is a compact eightstage device having low thermal mass compared with impactors such as the ACI. The use of absorbent glass fiber filter substrates increases its capacity to collect droplets without overloading. This miniaturized CI uses slot, rather than circular nozzles for stages 16, so that their sharpness of cut is slightly reduced (GSD values of most stages are in the range of 1.31.4). However, the presence of fewer stages compared with the ACI reduces overlap except for the first three stages, where some starvation is apparent from published calibration data.77 The impactor was developed for workplace personal exposure monitoring of particulates, and like the ACI, has subsequently been adapted for inhaler testing. As such, there are some concerns that are emerging with increasing user experience in the pharmaceutical industry.76 In particular, the acquisition and fitting of the specialized filters is difficult, and filter fibers can contaminate API work-up for assay. The analytical sensitivity of the technique is quite low, given the fact that only just over 13% of the aerosol leaving the nebulizer can be sampled in accordance with the methodology given in the CEN standard. The sampling configuration is a straightforward 22-mm internal diameter T-connector arrangement, but its particle sizeflow rate sampling characteristics have not been established. Finally, the use of sodium fluoride solution as a surrogate for drug formulations, whilst fulfilling the intent of providing baseline data with which to compare different nebulizer brands,79 is unlikely to be predictive of the behavior of suspension formulations where droplets can be formed that do not contain API.72 The low flow impactor technique has not been included in either the European or US compendia at the present time, although the European Respiratory Society has endorsed the test methodology in the CEN standard as a means of providing clinicians with reliable data for the wide variety of nebulizers that are in the marketplace. 80 The total flow rate of 15 L/min was chosen to match the mid-point flow of the sinusoidal pattern proposed for breathing simulator-based testing of these devices.79 However, there is some flexibility in the choice of this value, since the methodology in the CEN standard is claimed to be adaptable to virtually any breathing pattern.79 Recently, Jauernig et al., 76 based on the out-

352

MITCHELL AND NAGEL

come of preliminary experimental studies, suggested that the NGI operated at 15 L/min may be adaptable for characterizing nebulizer-produced aerosols following an adapted CEN protocol, since this CI can sample the entire aerosol. They noted that the simplest approach would be to utilize the existing design, calibrating the archival NGI at this flow rate to establish the correct stage cut sizes. However, as an alternative strategy, they showed that by plugging 50% of the nozzles for each stage, they could achieve comparable droplet size distribution data from a modified e-Flow nebulizer, operating the NGI at 15 L/min. They used the same stage cut sizes as those proposed for the unmodified NGI at 30 L/min (note that they reduced the area of the single nozzle for stage 1 by 50%). Although these configurations show promise for making reliable CI-based measurements of nebulizer-produced droplets, either option will require extensive validation studies, given that this impactor is being used in an application for which it was not specifically intended.

CI AUTOMATION
CI measurements are labor intensive, so that efforts have been under way in recent years to automate their use, particularly for inhaler development and batch release testing, where a large number of similar measurements must be performed on a regular basis. If CIs listed in the compendia are considered, both the ACI and MSLI are problematical to service using robotics, since they are both vertical stacks, requiring individual stages to be accessed without disturbing the others, and the collection plates tightly fit their holders (typically ,0.1 mm gap). Despite these difficulties, Smith 81 described a means of overcoming the handling problem by modifying the outside of the impactor with grooves and location balls, enabling the robot to locate and manipulate the individual components accurately. In the same article, he presented comparable size distribution data from an (unspecified) pMDI obtained by the automated ACI with results acquired manually, indicating that 3040 inhalers could be tested in a 24-h period by the automated system. The NGI by virtue of its horizontal stage layout, including the use of the MOC in place of a filter, lends itself more readily to automation. Although Smith has since pro-

posed an outline scheme for undertaking this task,82 a design for a fully automated system has yet to appear. A group of automation experts from member companies involved with the development of the NGI is also currently working towards the same goal.83 Fully automated CIs are highly expensive, an estimate for the automated ACI being close to US$1M. 81 As a means of controlling equipment cost, Miller et al.84 recently proposed simplifying the liquid handling procedures involved with API recovery from CIs by making the primary motion of the process in the liquid streams required for API dissolution and cleaning (so-called service-head approach), and applying the concept in relation to the NGI. Such a methodology avoids the need to move the components of the CI physically, resulting in a greatly simplified and therefore less costly approach to automation. However, for most users, lower cost aids that speed up the process of making CI measurements in a semi-automated way, for instance by applying stage coatings in a reproducible manner or by assisting with API recovery from individual stages, are likely to be more attractive than fully automated systems. Initial findings with one such system indicated both excellent API recovery (ca. 99% of label claim) together with greatly reduced cycle times compared with manual ACI operation.85

ADAPTATIONS TO CIs FOR USE WITH INHALERS: INDUCTION PORTS


Some form of inlet is required to ensure that the aerosol produced by the inhaler is sampled in a reproducible manner. From a practical perspective, most inhalers deliver their medication in the horizontal plane and entries to CIs are in the vertical downwards direction. The solution to the problem is the induction port, which also serves the purpose of mimicking to a greater or lesser extent depending on its design, the human oropharyngeal region. There are many designs of induction port (seven examples are given in Fig. 3), reflecting differing viewpoints on how inhaler aerosols should be sampled, 86 but by far the most commonly encountered is the metal right-angle bend described in both U.S. and European and Pharmacopeias. 1,2 In its role as a model of the entrance to the respiratory tract, the induction port collects almost all of the fast moving and so-called

CASCADE IMPACTORS AND SIZE OF AEROSOLS FROM INHALERS

353

FIG. 3. Various designs of induction port.86 (Reprinted from J. Aerosol Med. 11(S1), Dolovich, M., and R. Rhem. Impact of oropharyngeal deposition on inhaled dose, 112115, Copyright 1998, Used with permission of Mary Ann Liebert Inc.)

ballistic component of pMDI-produced aerosols formed by flash evaporation of the propellant, and therefore likely to deposit in the oropharynx.87 It also serves to remove larger and often aggregated particles generated by most DPIs upon inhalation. The USP/EP design is intended to provide a common benchmark to compare different formulations by standardizing its critical dimensions (internal diameter and unobstructed path length). Unfortunately, unlike an impactor stage, its collection efficiency is not easily determined, since flow is non-laminar at flow rates encountered typically for inhaler testing, and particles in the ballistic component have velocities greater than that of the surrounding airflow when entering the induction port. In an attempt to develop a better understanding of the processes that influence particle collection in the induction port, Stein and Gabrio 88 investigated quantitatively the deposition of solution-based pMDI HFA formulations at various locations within the USP/EP design as a function of flow

rate within the exceptionally wide range of 590 L/min. They discovered interestingly that inertial impaction is not the dominant particle deposition process; since turbulent deposition is more important, especially at the higher flow rates where the turbulent intensity of the flow moving through the induction port is at its greatest. Van Oort and Downey89 have argued that, in the case of pMDIs, blow-back of aerosol towards the inhaler can be exacerbated by the small size of the USP/EP design. In support, Van Oort and Truman90 have since shown that deposition of pMDI-produced albuterol within this induction port, expressed as a percentage of the label claim dose, decreases in a linear relationship with increasing length of the entry. A configuration based on the larger volume Twin Impinger inlet, manufactured out of either metal or glass was proposed as an alternative to the USP/EP design,89 based on experimental studies indicating that the proportion of the emitted dose contained in fine particles with dae , 5 mm entering the im-

354

MITCHELL AND NAGEL

pactor was increased. Earlier work by this group with pMDI-produced aerosols had indicated a trend towards increased fine particle transmission to the impactor with increasing induction port internal volume up to 5 L in capacity, with most improvement evident between 250 cm3 and 1 L.91 The underlying mechanism is believed to be the increased time for solvent/propellant evaporation to be completed within the larger volume induction ports. The 1L glass entry port developed recently by Sequeira et al.92 therefore probably represents a near optimum condition in terms of allowing the aerosol plume from the pMDI sufficient room to expand, and thereby enhancing fine particle delivery to the impactor. In their justification for such a large induction port, these authors claimed that the additional dose collected in the impactor affords greater sensitivity when following subtle changes in particle size distribution from one batch to another. This size of induction port should be viewed as an aid to enhance product quality control, rather than an attempt to mimic the geometry of the upper respiratory tract as a means of predicting lung deposition, given the relatively small capacity of the human oropharyngeal cavity. The SRI induction port developed by Williams and Witham,93 and originally proposed by Byron94 as an alternative to the USP/EP induction port, was also an attempt at an improved design of CI entry, primarily for quality testing low resistance DPIs at flow rates up to 120 L/min. The design intent was to enable an ACI to be operated at its standard flow rate (28.3 L/min) by introducing a flow of clean make-up air in an extended vertical section, followed by an isokinetic sampling arrangement to allow flow to enter the impactor without biasing the size distribution. Although the underlying principles appeared to be sound, concerns were subsequently expressed about the ability to sample the bolus of aerosol released from a DPI on actuation isokinetically and the inherent variability associated with the turbulent mixing process involving the make-up air.95 The Williams and Witham inlet has not become adopted as a standard induction port, since current compendial methods for DPI testing1,2 requires that a fixed volume (4.0 6 0.2 L) of air be sampled from the inhaler at a flow rate determined by the resistance of the inhaler. The additional dead volume introduced by this design is undesirable when the volume of air that can be sampled is restricted, since it affects the charac-

teristic flow rate-time curve for the inhaler. The move to variable flow rate testing, at least for DPIs, was spurred on by the development of impactors that were designed from the outset to be operated at a wide range of flow rates.51,52 In addition, the current method avoids concerns about sampling, since the entire dose emitted from the inhaler is collected. The testing of breath-actuated inhalers poses a problem with conventional induction port designs, since the act of connecting the inhaler to the inlet may trigger the breath-actuation feature. In response, Brouet et al.96 recently proposed a modified USP/EP induction port that incorporates a fast (540 msec) switching solenoid valve located in the horizontal section. In use, the inhaler is connected to the entry of the induction port with the valve initially switched to permit air to enter the impactor via a by-pass port. Flow is diverted to enter the induction port via the inhaler at the appropriate time. In addition to the challenges of breath-actuated inhaler testing, the evaluation of add-on devices for pMDIs, particularly HCs, also creates difficulty when simulating poor patient coordination, including mistiming of the inhalation maneuver and breath-holding following inhaler actuation. In an attempt to address this concern, Mitchell and Nagel 42 reported the use of an electro-mechanically operated shutter that attaches to the induction port entry. This apparatus can be used to simulate delay times from fractions of a second to several seconds in duration between actuation of an inhaler and the onset of sampling. It avoids modifications to the induction port itself, and adds less than 5 mL of dead volume to the sampling arrangement. In principle, the delay apparatus could be used with almost any induction port, although it has only been evaluated with the USP/EP design to date. In the context of a meeting of experts whose purpose was the development of an industrywide consensus towards more meaningful laboratory measurements for evaluating inhaler-produced aerosols, Dolovich and Rhem 86 noted that in vitro sizing with the various induction ports that are available may introduce inconsistencies with in vivo data from various laboratories. They therefore proposed the adoption of an inlet design that preferably most closely resembles the aerosol collection properties of the human mouth, pharynx and trachea. The USP/EP induction port, although widely adopted, does not meet this

CASCADE IMPACTORS AND SIZE OF AEROSOLS FROM INHALERS

355

criterion, as its internal geometry is too simple. Nevertheless, it could be argued that for a pharmaceutical manufacturer, the acquisition of consistent in vitro data from one batch to the next of a given formulation with a given design of induction port is more important than achieving comparability of data obtained with different designs. Under these circumstances, the choice of induction port (whether USP/EP or another type having simplified geometry) should be made to optimize the sensitivity of the CI method for the formulation being evaluated, as proposed by Sequeira et al.92 On the other hand, a number of groups have in the last few years recognized the over-riding importance of establishing reliable in vivo correlations with laboratory-generated particle size distribution measurements in the context of evaluating likely clinical performance.97 This understanding has resulted in the creation of several induction port geometries based on both cadaver casts and MRI imaging of live patients, in order to model as closely as possible the clinical situation.98,99 The effect of induction port choice on the dose sampled by the CI can be significant. For instance, Berg99 observed that the fine particle dose (dae , 5 mm) from a budesonide pMDI delivered by a large volume holding chamber (HC) to an ACI operated at 28.3 L/min with a glass Twin Impinger induction port was comparable at 35% label claim dose (200 mg) to that from the pMDI alone. However, when an induction port whose entry profile was based on the anatomical model of an adult throat replaced this inlet, fine particle dose ex HC doubled from 20% to 40% of the label claim dose compared with that from the pMDI alone. This difference was more pronounced (8%, pMDI alone; 25%, pMDI 1 HC) when the induction port was changed to one based on an anatomical child throat. At about the same time, Olsson et al., 100 using the charcoal block technique, compared the lung deposition of albuterol in healthy adult volunteers with CImeasured fine particle dose (dae , 5 mm) for three dry powder inhalers (DPIs) and a pMDI. A better in vivo/in vitro correlation was achieved with data obtained using an adult throat replica compared with a glass Twin Impinger induction port, the latter significantly over-predicting the lung deposition that was actually achieved. Their use of Brij-35 (polyoxyethylene 23 lauryl ether) surfactant (0.75 g) in glycerol (25 g) applied in ethanolic solution (7 mL) as a coating to the replica throat is an interesting extension of the

idea of more closely modeling the in vivo situation by simulating the wet mucosa of the upper respiratory tract. More recent studies by Berg et al.101 with induction ports modeling five different human oropharyngeal anatomies, including adult with tongue up or down, mouth open or mostly closed, as well as that of a ,3.5-yearold child, have demonstrated as much as a fivefold difference in fine particle delivery for the delivery of CFC- and HFA-formulated fluticasone propionate and budesonide. These differences were attributed to the varying geometry of the inlets and were also found to be formulation independent. Srichana et al.102 reported an investigation into the applicability of replacing the TI induction port with a cast of an adult male throat having wetted interior surfaces, coming to the conclusion that the cast had higher retention when sampling lactose carrier-based DPI-generated aerosols at 60 L/min. Given what has already been presented, it is unsurprising that Massoud et al.98 observed significantly greater retention of DPI-generated aerosols sampled by small compared with large size oropharyngeal models made from MRI scans of patients, representing extreme dimensions likely to be encountered in clinical practice. Alternatives to an induction port as CI entry are possible in the context of nebulizer testing, since these inhalers do not generate ballistic droplets and there is no need to simulate bolus delivery of medication, as is the case with DPI testing. For instance, the use of a 22-mm diameter T-connector constructed from components used in mechanical ventilator circuitry has been proposed to enable flow from a jet nebulizer to be sampled by a CI operating at a fixed flow rate, whilst the air flow leaving the nebulizer is varied to simulate tidal breathing. 103 Although not developed as an induction port through a formalized design process, this readily available arrangement having reproducible internal geometry may be particularly useful for the testing of breath-enhanced or breath-actuated nebulizers, where breath-simulation is essential to get the air entrainment and/or breath-actuation feature to function as they would under patient use. Like most induction port designs, the aerodynamic characteristics of this T-connector inlet have not yet been evaluated either by computational fluid dynamics or experimentally with monodisperse particles. Furthermore, the CI did not sample the entire aerosol produced by the nebulizer.

356

MITCHELL AND NAGEL

Nevertheless, indirect evidence of its suitability is given in an article in this issue,104 which demonstrates a correlation between lung deposition based on g-scintigraphy and CI-measured FPF,2.5 mm . These in vitro measurements were made sampling jet nebulizer-produced droplets at a very low flow rate (1 L/min) into a 10-stage QCM CI. Deposition within the T-connector is believed to represent the high inertia and therefore coarse particle component of the nebulized aerosol that is most likely to be deposited in the upper respiratory tract. The challenge of testing oral inhalers in a realistic manner, but at the same time preserving simplicity in technique has been addressed in a different way by Miller and Purrington.105 They described the comparative in vitro evaluation of the metal induction port for the Marple-Miller (MMI) series of impactors with a cast adult throat replica, utilizing the aerosol penetration characteristics of a variety of undisclosed CFC- and HFA-formulated solution and suspension pMDIs. The portion of the emitted dose that entered the CI via the MMI induction port was observed to follow closely the equivalent measure of aerosol penetration through the replica throat, sampling at both 30 and 60 L/min. In an additional comparison with the MMI and USP/EP induction ports, they discovered significant differences with some formulations, especially at the lower flow rate, with greater penetration of the emitted dose from the inhaler generally, but not always apparent with the USP/EP induction port. Although such data support the use of the MMI induction port as a surrogate at least for adult throat geometry when evaluating pMDIproduced aerosols, further studies of this sort are needed to establish the comparability of this inlet, or a similar shaped but smaller induction port with anatomically correct child throat models. In passing, it is self-evident that the use of an induction port would be inappropriate for CI measurements to characterize inhalers used with devices that instill the aerosol directly to the region of the carina via an endotracheal tube (ET), bypassing the oropharyngeal region. In this context, Mitchell et al.106 described the use of a standard ACI to evaluate the performance of a holding chamber fitted with an 8.0-mm internal diameter ET, intended for mechanically ventilated adult patients. The aerosol leaving the ET was sampled on-axis directly into the impactor inlet cone, enabling the size distribution of the

API to be determined at the point of delivery to the lower respiratory tract. A further refinement was the incorporation of humidification and heating of the air supply passing through the holding chamber to the impactor to simulate more closely the conditions in a mechanical ventilator circuit.

PRE-SEPARATORS
A pre-separator is often required when sampling aerosols produced from DPIs, since these formulations in many instances contain the API attached to the surface of much larger glucose or lactose carrier particles. The shear forces generated by inhalation detach some, but not all of the API particles from the carrier material, and the former are sufficiently fine to penetrate beyond the upper respiratory tract into the lungs.28 Both U.S. and European compendia1,2 refer to the use of a pre-separator for DPI-based particle size distribution measurements with the ACI, recommending that its interior surfaces be coated either with a tacky agent in the same way as the collection surfaces of the CI stages, or with up to 10mL of a suitable solvent to eliminate particle bounce and re-entrainment. Ideally, the pre-separator should not starve the first stage of the impactor of particles. However, the pre-separator used with the ACI has been shown by calibration to have its cut size close to 9 mm in aerodynamic diameter, almost identical with that of the first stage (stage 0) at 28.3 L/min.48 Its sharpness of cut is relatively poor at this flow rate by comparison to equivalent values for the CI stages (GSDpre-sep is about 1.548 ), due primarily to the influence of gravity. The pre-separator therefore starves the second as well as the first stage of the impactor. Sethuraman and Hickey107 have studied the size-separating performance of the ACI pre-separator at 60 L/min using computational fluid dynamics, supported by experimental work in which polydisperse fluorescent particles in the range of 45125 mm were sampled by this impactor. Their modeling revealed the existence of low velocity locations, located particularly above the entry orifices, where particles can deposit away from their intended sites, thereby increasing wall losses. Interference in the flow between adjacent nozzles was cited as a cause of its poor size-selectivity. Their experiments indicated that coating the internal surfaces with silicone oil (un-

CASCADE IMPACTORS AND SIZE OF AEROSOLS FROM INHALERS

357

specified viscosity) improved the size-selectivity (sharpness of cut) of the pre-separator, reducing the carry-over of large particles to the impactor. Unfortunately, this group did not calibrate their pre-separator using monodisperse particles, so that GSD pre-sep cannot not be established. However, apart from improvements in size-selectivity brought about by coating the interior surfaces, GSD pre-sep would be expected to fall below 1.5 with increasing flow rate, given the increased dominance of inertia in the size-separation process at higher Ref. Hence the overlap between its collection efficiency curve and those of the uppermost impactor stages would be expected to be diminished at flow rates 5 60 L/min, compared with that observed at 28.3 L/min. Recognizing the limitations of the ACI pre-separator, the consortium developing the NGI designed its pre-separator to function as a single component, but with two distinct steps in the size-separation process.108 The incoming aerosol is first passed through a so-called scalper impingement stage that removes the coarsest particles. The remaining aerosol then passes immediately through a more conventional impaction stage before leaving the pre-separator. Calibration with monodisperse particles has demonstrated that its GSD pre-sep is close to 1.3 at 30, 60, and 100 L/min,57 which is comparable with values of GSD stage for well-designed impactors. Its measured d50 values (10.0, 12.7, and 14.9 mm at 100, 60, and 30 L/min, respectively) are sufficiently separated from the corresponding values for stage 1 (6.07, 8.29, and 11.4 mm) that starvation of the first stage of the impactor should not occur to a significant extent.

ADAPTATIONS OF CIs WITH BREATHING SIMULATORS


Although CIs must operate at constant flow rate for the stage cut sizes to remain stable during a measurement, there have been several attempts to link these instruments directly with breathing simulators in order to arrive at aerosol transport conditions that more closely mirror the clinical situation. In the case of DPI testing by the compendial procedure,1,2 the inhalation portion of a single breathing cycle is simulated by opening a twoway solenoid valve located between the CI and pump for a pre-determined time. The flow en-

tering the CI can take a short but significant time to reach the nominal flow rate for the test, depending on the magnitude of the dead volume in the system, during which the stage cut sizes rapidly decrease to their final stable values. It is normal, however, to treat the measurement as having been undertaken at the nominal flow rate throughout the entire sampling period, so that fixed cut sizes, such as those given in Tables 2 and 3, can be used to determine the size distribution data. In the case of pMDI testing, it is normal to couple the inhaler mouthpiece directly to the induction port of the CI with the flow set at the nominal value, actuating at the appropriate time. Under these circumstances, the assumption of stable cut sizes throughout the measurement is valid, provided that flow rate control is adequate (typically 65% of nominal1,2 ). The testing of HCs poses some additional challenges, since these devices typically have inhalation and exhalation valves whose opening and closing characteristics are flow rate dependent.42 Attempts thus far to interface CIs directly to breathing simulators for testing these devices have proved somewhat problematic. Fink and Dhand109 described an arrangement in which a QCM impactor samples aerosol emitted from an HC into a 1-L plenum that is also connected to a ventilator circuit. Although simple in concept, this configuration has the drawback that the flow velocity profile at the inlet to the CI is constantly changing as a result of the imposed breathing pattern. Sampling of polydisperse aerosols typically produced by inhalers should ideally be isokinetic (i.e., the air velocity at the inlet should match that of the air flow from which the sample is being taken) to eliminate size-related bias.5 The systems developed by the group at the University of Alberta110112 and by Foss and Keppel113 are more sophisticated, since they both attempted to separate the continually varying flow generated by the breathing simulator and fed to the HC from the constant flow path required by the CI. In the Finlay and Zuberbuhler model110112 (Fig. 4), an ACI operating at 28.3 L/min was allowed to sample the aerosol emitted from the HC fitted to a replica face. The signal from a stepper motor-controlled piston-driven breathing simulator was used to actuate a two-way solenoid valve, pushing sufficient flow of air into the system during the inhalation portion of each breathing cycle to enable the impactor to operate normally and allow the system to inhale from the holding cham-

358

MITCHELL AND NAGEL

FIG. 4. Breathing simulatorcascade impactor system developed by Finlay,110 and Finlay and Zuberbuhler. 111,112 (Reprinted from Int. J. Pharm. 168, Finlay, W.H., Inertial sizing of aerosol inhaled during pediatric tidal breathing from an MDI with attached holding chamber, 147152, Copyright 1988, with permission from Elsevier.)

ber. At exhalation, the solenoid valve closed, so that the piston pushed flow back towards the CI and HC, correctly simulating the exhalation flow profile at the replica face. At the same time the CI was able to continue to draw a constant flow without evacuating the system. This arrangement was used to study the behavior of several HCs at low flow rates of ,10 L/min, corresponding to use by small children.110112 However, in some circumstances, particularly when simulating use at low flow rates, it may be possible for pressure pulsations arising from the operation of the fastacting solenoid valve in this arrangement to pop open an otherwise stuck closed inhalation valve with at least one type of HC. This suggestion is based on the outcome of testing of HCs elsewhere (Vent-170, SpaceChamber). 114 Both were shown to have a tendency for valve adhesion preventing delivery of medication at low flow rates, when tested using a breathing simulator without CI, and employing filter collection to capture the total emitted dose at low flow rates (tidal volume , 100 mL). The system developed by Foss and Keppel 113 both avoided pressure pulsations and generated a sinusoidal breathing pattern, which could also be readily adapted to generate other waveforms. In their arrangement, a supply of pressurized air was fed continuously to a Y-shaped connector, the other arm of

which was connected to a breathing simulator (Fig. 5). The combined flow entered a T-connector, one arm of which was coupled to the HC and the other arm that was connected to the CI. Aerosol from the HC passed through a USP/EP induction port at variable flow rate, simulating the action of inhalation. After passing the straight section of the T-connector, the aerosol stream attained constant velocity under the influence of the vacuum pump that is connected to the CI. This arrangement requires careful flow control to avoid losing aerosol in transit from the holding chamber to the impactor, and is also limited to inhalation flow rates that are less than the flow rate required by the CI. However, the timing of the breathing cycle could be adjusted to simulate actuation of the pMDI both in and out of phase with the onset of inhalation to simulate realistic use poorly coordinated patients, for which HCs are frequently prescribed. The Electronic Lung developed by GlaxoSmithKline Plc was an attempt to allow a more accurate simulation of DPI performance during patient use than that which can be achieved with direct constant flow rate sampling methods. In this apparatus, aerosol from a DPI was first inhaled into a 6 L capacity vessel,115 by withdrawing a piston through a chamber connected to the chamber via a sidearm at its base (Fig. 6).

CASCADE IMPACTORS AND SIZE OF AEROSOLS FROM INHALERS

359

FIG. 5. Breathing simulatorcascade impactor system developed by Foss and Keppel.113 (Reprinted from Respir. Care. 44(12), Foss, S.A., and J.W. Keppel. In vitro testing of MDI spacers: a technique for measuring respirable dose output with actuation in-phase or out-of-phase with inhalation, 14741485, Copyright 1999, with permission from Daedelus Enterprises Inc.)

A feedback mechanism ensured that the actual pressure drop across the DPI corresponded to the demanded pressure drop associated with the inhalation maneuver being simulated, and the piston movement adjusted accordingly. The aerosol within the chamber was therefore claimed to be representative of that inhaled in normal use. Following inhalation, a solenoid valve was opened

so that the aerosol could be sampled at 28.3 L/min via a standard ACI connected to the base of the chamber for sufficient time to ensure complete emptying of the chamber with two complete air changes. Although this system has the advantage that patient-derived inhalation flow ratetime profiles can be used to simulate actual DPI use, it is vulnerable to size-related sampling bias,

FIG. 6. Electronic lung.115,116 (Reprinted from J. Aerosol Sci. 29(8), Burnell, P.K.P., A. Malton, K. Reavill, and M.H.E. Ball. Design, validation and initial testing of the Electronic Lung Device, 10111025, Copyright 1998, with permission from Elsevier.)

360

MITCHELL AND NAGEL

principally caused by gravitational sedimentation within the chamber. However, validation experiments using polydisperse particles of comparable size range to those produced by commercially available DPIs, indicated that particles finer than about 5.8 mm in aerodynamic diameter are sampled with high efficiency (.80%).116 Although it is well suited for DPI evaluations, the Electronic Lung is not readily applicable to the testing of HCs, where it may be necessary to simulate exhalation through the device in the context of evaluating performance with an uncoordinated patient. In the context of DPI testing, Finlay and Gehmlich 117 recently adapted the Finlay-Zuberbuhler system to permit low to moderate resistance DPIs (Spiros and Ventodisk [Diskhaler]) to be tested at a wider range of inhalation flow profiles simulating actual patient use (average inhalation flow rates varied from 21 to 162 L/min depending on inhaler type). At the same time, their system permitted a CI to sample the aerosol via an anatomically correct oropharyngeal cast at constant flow rate (Fig. 7). They were able to demonstrate the important finding that for these inhalers at least, CI-derived size distributions measured directly at

constant flow rate (via the oropharyngeal model) were comparable with those obtained using both fast and slow breathing patterns with their breathing simulator-CI system, as long as the flow rate was carefully chosen to be comparable with those obtained in the simulations. In order to encompass the wide range of flow rates encountered with these DPIs, it is worth noting that Finlay and Gehmlich used a two-stage virtual impactor to permit sampling via a CI at 30 L/min from the air stream leaving the inhaler through a connecting tube to which make-up air could be added to achieve a total flow rate of 285 L/min. This system probably represents the state-of-the art sampling arrangement for obtaining particle size data from DPIs being operated under realistic conditions of use, but it remains to be evaluated with higher resistance inhalers. Given the outcome of the Gehmlich and Finlay study with DPIs,117 as well as evidence that, for at least one press-and-breathe pMDI-based formulation (HFA-fluticasone propionate), flow rate variations between 28.3 and 60 L/min had an insignificant impact on CI-measured size distribution data,118 a case might be made for simplifying in vitro testing under realistic breathing

FIG. 7. High flow rate virtual impactorCI system developed by Finlay and Gehmlich. 117 (Reprinted from Int. J. Pharm. 210, Finlay, W.H., and M.G. Gehmlich, M.G. Inertial sizing of aerosol inhaled from two dry powder inhalers with realistic breath patterns versus constant flow rates, 8395, Copyright 2000, with permission from Elsevier.)

CASCADE IMPACTORS AND SIZE OF AEROSOLS FROM INHALERS

361

conditions to the assessment of TED by filter collection. A CI could then be used as an adjunct procedure, testing at constant flow rates judiciously chosen to be appropriate for the intended patient user groups. However, it is recognized that further research is needed to ascertain how applicable such a simplified approach would be to the testing of the wide variety of inhalers that are available. On a precautionary note, the data from Smith et al.119 would indicate that although such a simplification might be justifiable with conventional press-and-breathe and breath-actuated pMDIs, it may not be appropriate with certain DPI-delivered aerosols.

USE OF CIs WITH INHALERS FOR NASAL DRUG DELIVERY


Nasal drug delivery is rapidly evolving into a popular route for administering both topical and systemic pharmaceutical agents, resulting in recently issued FDA draft Guidances for Industry covering appropriate in vitro assessments35 and bioavailability/ bioequivalence issues.120 In both documents, even though the API is not directly quantified, light scattering techniques such as laser diffractometry121,122 are recommended for aqueous nasal products using mechanical pumps,120 since this class of inhaler typically produce droplets with dp in the 20 to 200 mm range,122 which is well beyond the region in which inertial size-separation methods are effective.5,9 Despite this limitation, CIs are also proposed for the characterization of API contained in fine droplets, but mainly to provide assurance that the mass of drug in droplets with dae of #10 mm is comparable between products.123 There is interest in quantifying this so-called fine component, which is typically ,5% of the dose, since such droplets can pass beyond the nasal cavity leading to undesired deposition elsewhere in the respiratory tract.124 However, this type of measurement has low sensitivity for reasons already discussed.125 Suman et al.11 used a nasal cavity model made from an inverted twin-necked round-bottom flask, and improved measurement precision by using a shortened ACI (pre-separator and impaction stages 0, 1, and 2 and filter only) thereby increasing analytical sensitivity for the API. They actuated the nasal spray upwards into the sphere via one neck of the flask, sampling vertically

downwards from the other neck located at its base at 28.3 L/min. They were able to demonstrate comparable fine component delivery from two types of commercially available aqueous nasal spray pump devices, in that 97.8 6 2.1% and 99.8 6 3.8% of the dose from nicotine solution sprays delivered by either device was contained in particles with dae of $9 mm. More recently, Guo et al.126 extended this technique to illustrate how the volume of the glass entry port, which was varied from 300 mL to 5 L in capacity, affected the magnitude of the fine component (,3.8% of the dose ex actuator under all conditions) collected by their short stack ACI. Although no effect was evident at 30 L/min, reducing the volume of the sphere slightly decreased the magnitude of the fine component of the dose. As might be expected, the fine component collected in the CI increased slightly when the flow rate was raised to 60 L/min. The authors noted that these data contrast with the outcome of an earlier in vivo study by one of the co-authors,127 in which no lung penetration was observed with a similar nasal spray, so the conclusion was drawn that an anatomically accurate nasal cavity model would probably result in even lower fine component with their in vitro measurements with these oversized entry ports. Doub and Adams128 recently reported an examination of the fine component of aqueous beclomethasone dipropionate (BDP) nasal spray pump aerosols using an ACI also equipped with either a 2- or 5-L spherical glass induction port. Their short stack CI, which was constructed with components normally used at 90 L/min (pre-separator, stages 22, 21, 5, and filter), but operated at 28.3 L/min, was slightly more sensitive than the configuration used in the studies by Suman et al.11 and Guo et al., 126 since the cut size for stage 22 is 13.6 mm, compared with 9.0 mm for stage 0 at this flow rate, so that more of the dose could enter the CI. However, Doub and Adams128 acknowledged that the absolute mass of API per actuation entering the CI was still a very small fraction of the dose ex actuator (0.51%). With the advent of the NGI, there may be an opportunity to re-evaluate the precision of this measurement, given the good sharpness of cut for its pre-separator (cut size 5 12.7 mm at 60 L/min).57 This component might conceivably be inverted and used as a substitute entry port. Furthermore, the NGI can be made to function effectively with a reduced number of operating stages, by inserting

362

MITCHELL AND NAGEL

deep collection cups in place of the normal cups for those stages where particle collection is not required, thereby optimizing the analytical sensitivity.83 The CI is a more appropriate technique for sizecharacterizing the aerosols produced from propellant-driven nasal metered-dose inhalers (socalled n-pMDIs), since these aerosols comprise finer droplets produced as a result of the energy available from propellant expansion when the inhaler is actuated. However, as with all inhaler testing, selection of the most suitable CI configuration is determined by both the geometry of the induction port (in this instance to accommodate upward actuation of the inhaler), as well as the cut sizes for the impactor stages at the chosen flow rate. Yu et al.61 first reported the use of a six-stage Delron CI equipped with 5-L glass chamber entry made from a round-bottomed flask to simulate the nasal cavity that collected droplets with dae . 16 mm from a mechanical pump nasal spray. More recently, Ostrander et al.129 described the use of a more realistically shaped glass induction port (without turbinate structures) that fitted above a Twin-Impinger induction port having its entry (mouth) stopped to simulate a closed mouth. This arrangement therefore simulated both the human nasal and oral cavities as an entry to their CI. Leung et al.130 questioned the value of CI-based measurements for this class of inhaler if the entry to the CI only mimics a simplified geometry of the nasal passageways. Their ACI-based data, obtained at 28.3 L/min with glass and silicone rubber induction ports shaped to mimic human nasal anatomy, showed the presence of a significant non-ballistic fraction having MMAD values of 2.64.2 mm, depending upon the nature of the propellant. However, good nasal deposition was also achieved in their nasal cavities, particularly the one manufactured from silicone rubber containing turbinate-like structures within its confines. Given the complexity of the nasal passageways, in particular the importance of turbulence as a means of depositing fine particles, which is not well reproduced by conventional induction port design, they concluded that in vivo nasal deposition from n-pMDIs is likely to be much higher than that indicated using a smooth-surfaced glass induction port that can only represent the overall shape of the nasal cavity. Further work is needed to resolve this issue, which will hopefully lead to the development of appropriate induction port-CI configurations for use with n-pMDIs.

OTHER CONCERNS RELATING SPECIFICALLY TO CI OPERATION INTERNAL (INTER-STAGE) LOSSES


Internal losses of particles on CI surfaces other than those intended for collection (so-called wall, or inter-stage losses) are related to the issue of particle bounce, since particles with tacky surfaces are more susceptible by virtue of their increased ability to adhere to all types of surfaces in which they come into contact.9 An upper limit of 5% of the total delivered drug mass per actuation from the inhaler has been defined as a CI system suitability requirement in the U.S. Pharmacopeia, 1 to limit the impact of internal losses on measurement accuracy. However, a recommendation is also made that, in the event that the losses exceed this limit, the procedure should be performed in such a way as to include wall losses in the assessment of API. In practice, it is not possible to apportion such losses from stage to stage within the CI, because they lack size classification. The alternative option to utilize a different type of CI is therefore to be preferred. Inter-stage losses are also particle size dependent, particularly when dae . 10 mm, as illustrated by calibration studies in which different sizes of uniform particles were sampled by ACIs operated at 28.3 L/min.48,55 It follows that the pre-separator is the most vulnerable component of the CI to such losses, but little has been done to quantify them for the commonly encountered CI designs. This is probably because for most inhaler testing, the pre-separator is not considered as an additional impaction stage, but as a pre-classifier of particles whose inertia is greater than that associated with the cut size of the first impaction stage, and therefore unlikely to reach the lung. Under this assumption, losses are only important if they are related to API that cannot be recovered or are associated with fine particles that should have entered the impactor. In this context, Sethuraman and Hickey 107 identified low-velocity regions above and after exiting the exit orifices in an ACI pre-separator, in which increased internal losses would be expected to occur away from the intended collection area on the floor below the entrance orifice, and which might therefore not be recovered. Unfortunately, the magnitude of these losses could not be quantified by their experimental technique that involved polydisperse DPI-generated aerosols, and that was more focused on the use of a coating to eliminate

CASCADE IMPACTORS AND SIZE OF AEROSOLS FROM INHALERS

363

particle deaggregation. Data from an investigation using monodisperse particles of appropriate sizes would therefore be valuable in determining the magnitude of this non-ideal collection behavior as a function of particle size as well as flow rate. However, on a cautionary note, the surface properties of the calibration particles will affect their adhesion, so it may be impossible to separate unambiguously true internal losses due to flow irregularities from those caused by particle bounce and relocation within the pre-separator. Although inter-stage deposition with CIs has a direct impact on the material recovery of API, data quantifying such losses are very limited in connection with inhaler testing. In an attempt to develop an empirical relationship linking such losses with particle size for one particular CI, Marple et al.21 expressed ACI-based internal loss data obtained at 28.3 L/min,48 in terms of an empirical relationship: L(dp ) 5 0.06(exp{0.1dp}) (10)

in which L represents the sum of all inter-stage losses within the CI. These losses exceeded 10% of the total mass entering the impactor for particles larger than 5 mm in aerodynamic diameter. More detailed analysis of their distribution as a function of particle size demonstrated that they clustered about the stage at which the particles should have been collected.48 Internal losses within a 150-series MMI operated at either 30 or 60 L/min were lower than those determined for the ACI, being ,5% for particles with dae in the range of 119 mm, and typically in the range of 12%, based on data obtained with liquid droplet (fluorescent-tagged oleic acid) aerosols.52 The lower losses for the MMI were attributed as much to the choice of calibration aerosol (increased retention of solid, but moist and therefore tacky methylene blue particles on internal surfaces in the ACI calibration), as to impactor design. However, if this is the case, it follows that the magnitude of internal losses within the CI, like those referred to for the pre-separator, will be formulation/inhaler dependent, and this assertion is supported by the outcome of a comparative assessment of four types of pMDI- and DPI-generated aerosols by Olsson et al.43 They reported that total inter-stage losses in an ACI were 1.6% of the inhaler-delivered dose for a budesonide pMDI, 2.4% for an albuterol DPI, 3.4% for a budesonide DPI and 3.5% for an ipratropium bromide DPI.

Equivalent losses within a MSLI were lower (,1.2%), but followed the same pattern. They also found that inter-stage losses in a model 160 MMI were similar to those in the ACI, with the exception of the albuterol DPI, where losses were significantly higher, at 9.6%. These anomalously high losses were attributed to overloading of the grease-coated surfaces of the upper stages in the MMI with excessive amounts of carrier lactose, resulting in over-sized particles cascading further down the CI, implying the need for a high-capacity pre-separator with this CI for this type of formulation. Apart from the isolated measurement just considered, these internal losses were all comfortably below the U.S. Pharmacopeia system suitability limit, so that variations between formulations/inhaler types were small in terms of the magnitude of the label claim dose. However, the authors cautioned that a detailed review of the underlying stage data revealed large relative differences between the distributions of losses between the combinations that were tested, counseling the need for validation with respect to each application (product-CI methodology). Internal loss data for production NGIs have yet to be published, but measurements made with prototype instruments indicate that they amount to ,5% of the inhaler-delivered dose from both pMDIs131 and DPIs,132 provided that the pre-separator is used with certain high unit dose pMDIdelivered formulations, and is appropriately coated with an agent to eliminate particle bounce.

OPERATOR CONTROLLABLE VARIABLES AFFECTING PSD ACCURACY


It is generally agreed that the CI measurement process is complex and labor intensive, but at the present time it provides the only way to determine inhaler particle size distributions by quantifying the mass of API separately from other components in the formulation. The PQRI working group developing methodology for identifying the causes of material balance failures have addressed in some detail the numerous operator-controllable variables that influence both the accuracy of this parameter as well as that of the size distribution measurement.39 Although issues such as choice of solvent and technique for API recovery are product dependent and therefore outside the scope of this

364

MITCHELL AND NAGEL

review, the choice whether or not to use a collection surface coating, as well as air leakage and flow rate control, are technique dependent, and are therefore discussed further.

COLLECTION SURFACE PREPARATION


Coating of CI stage surfaces with an agent to improve particle adhesion is well understood to be necessary to prevent bounce and re-entrainment of solid particles, 26,49,50,13 3 with the exception of the MSLI, where the liquid contained in each impaction stage fulfils this purpose, thereby avoiding the need for further treatment, other than to rinse all interior surfaces, including the walls and ceiling of each stage, diligently to recover the API quantitatively. 60 Particle bounce biases the measured size distribution data to finer sizes, and in the case of DPI-testing, some form of coating with a viscous, compliant agent, such as silicone oil as one of many examples, is almost essential.94,134 This precaution is necessary because of the high probability of elastic impact between the solid particles and the otherwise stiff collection surface, associated with a large coefficient of restitution.135 Coating of the interior surfaces of the ACI pre-separator has also been recommended as a means of reducing the break-up of large aggregate particles at impact from some DPI-based formulations. The pre-separator, therefore, collects aggregates that would not normally fragment upon inhalation, appropriately, and bias towards finer particle sizes in the size distribution measurements is avoided.107 Many pMDI-based formulations contain surfactant, which is compliant, and therefore more likely to result in inelastic collisions of drug-containing particles on impact with CI collection surfaces, thereby increasing the likelihood of effective adhesion. However, there is evidence that stage coating may be necessary, particularly for single- or two-actuation measurements of pMDI performance simulating guidance given in the patient insert,136 despite the fact that these inhalers are generally tested at lower flow rates than DPIs. For instance, Nasr et al.137 observed a so-called loading effect, whereby with uncoated surfaces in either an ACI or 150-series MMI operated at 28.3 and 30 L/min, respectively, one or four doses of pMDI-delivered albuterol particles were subject to significant bounce, whereas this phenomenon was not observed (to

a significant extent) when 10 actuations were delivered. By contrast, good particle adhesion was observed with either impactor when the stage collection surfaces were coated with glycerin or silicone fluid. Nasr and Allgire 138 and El-Araud et al.139 have also reported similar findings. These findings support the advice to treat coating of stages as an issue to be considered as part of the method validation process on a formulation/inhaler/CI basis.

LEAKAGE AND FLOW RATE CONTROL


The elimination of leakage and achievement of stable flow rate control for CI-based measurements are both factors that influence measurement accuracy, but the identification and removal of leaks has received scant attention in the literature. Since in almost all configurations, the airflow through the CI system carries the entire sample of particulate emitted from the inhaler, this stream must be directed in a controlled pathway. Leaks within the CI, for example caused by defective inter-stage seals in the case of the ACI, will provide an additional route for the in-flow of ambient air. Such leaks, including those at the connections between the inhaler and induction port as well as those associated with a pre-separator (if used) are not easily detected, and therefore must be guarded against by appropriate cleaning and maintenance. 140 Apart from the possibility of increasing local inter-stage deposition by deflecting the airflow from its proper pathway, leaks will reduce the upstream airflow, if flow rate is monitored downstream of the CI, decreasing flow velocity at the nozzles of the affected stages. Conversely, if flow rate is set correctly at the entry to the CI, leakages will increase the actual flow through stages that are downstream. Either situation will result in unpredictable changes to the cut sizes of the affected stages. Although leakages may have an ill-defined but possibly significant impact on the accuracy of the particle size distribution measurements, the material balance is unlikely to be affected, since there is no opportunity for material to escape from the system.39 Since the pressure differential within the CI and the surrounding ambient atmosphere is greatest at the stages nearest to the vacuum source, it follows that air leakage due to sealing defects are most likely to occur at these locations. At least one manufacturer has therefore developed a test fa-

CASCADE IMPACTORS AND SIZE OF AEROSOLS FROM INHALERS

365

cility in order that a user can verify that each stage of their CI is leak-tight before use.83 The size-discriminating ability of CIs is dependent on flow and hence particle velocity through the nozzles of each stage (equation 3). It therefore follows that the volumetric flow rate at the CI inlet is the parameter that must be established correctly.5 For testing pMDIs and nebulizers, this process can be relatively easily accomplished by the use of a soap-film flow-meter that has negligible resistance to flow, located upstream of the CI,140 since there is no requirement to establish a pressure drop at the inhaler interface, as is the case with DPI testing.1,2 A calibrated dry gas meter is likely to be more suitable for flow rates in excess of 30 L/min.141 However, if other types of flow-meter are used, the measured flow rate should be corrected to the conditions at the induction port entry.39 From the analysis of CI operating principles, it is evident that stage cut sizes will only be fixed if the volumetric flow rate through the CI is kept constant during the entire aerosol sampling procedure. This criterion is readily achieved when evaluating either pMDIs or nebulizers, as there is no need to combine the process of sampling by the CI with operation of these classes of inhaler. For pMDI testing, particularly when evaluating the impact of delayed inhalation through an HC, the practice of actuating the inhaler followed by applying the vacuum to permit the CI to sample142 should be avoided, as the significant internal airspace (dead volume) within the CI system, including induction port, pre-separator (if used), and pipe-work delays the rise in flow rate to the final value. During this period, which can be of the order of 1 sec, and which takes place when most of the aerosol has yet to be sampled, the stage cut-points are undefined, ultimately leading to bias that is difficult to quantify in the measured size distribution data.42 Flow rate setting to evaluate DPI performance can be difficult to accomplish when undertaken following compendial procedures, since it is necessary first to attach the DPI to the induction port of the CI, and adjust the flow rate until the pressure drop across the inhaler (measured at the induction port entry) is 4 kPa, chosen as being broadly representative of pressure drops produced by patients using DPIs.143 The DPI is then replaced by a flow-meter that is capable of providing the volumetric flow rate either directly or through an appropriate pressure correction. Pro-

viding that critical (i.e., sonic) flow is maintained at the regulating valve located between the CI and vacuum source, the assumption is made that exchanging the DPI for the flow-meter will not affect the flow rate at the CI inlet. Cox et al.144 identified two concerns with this methodology. Firstly, all currently available flow measuring devices introduce some disturbance into the flow being measured, making it difficult to select a suitable reference location to define exit flow rate (equivalent to CI entry flow rate). Although the critical orifice prevents pressure pulsations from the vacuum source being propagated upstream to the CI and inhaler, the mass flow rate through the system is affected by changes to air pressure and density upstream of the orifice. The authors conceded that the effect of such changes on CI inlet volumetric flow rate is negligible if: (1) The pressure drop across the critical orifice is much more than twice the upstream pressure drop across inhaler and CI system. (2) The pressure drop across both DPI and flowmeter are small compared with that across the CI (both likely conditions for most low and medium resistance DPI testing). They concluded by proposing that the compendial method be followed as far as setting the pressure drop at the induction port to 4 kPa with the inhaler attached to the induction port and establishing critical flow at the regulation valve downstream of the CI. The procedure was then modified by replacing the DPI with a flow-meter and second regulation valve, both located upstream of the induction port. This additional regulating valve was adjusted to restore the pressure drop at the impactor inlet to 4 kPa, thereby giving the same inlet flow rate to the CI as with the inhaler in place. Olsson and Asking145 more recently showed, by inserting variable flow resistances in place of an inhaler, that the compendial method which relies on achieving critical flow downstream of the CI to make the volumetric flow rate at its entry insensitive to changes in pressure drop at the inlet, is valid to a pressure drop of at least 12 kPa. This limit encompasses the range likely to be experienced in the testing of the highest resistance DPIs.146 Olsson and Asking145 further demonstrated that under these conditions, the volumetric flow rate downstream of the variable resistance is constant (varying mass flow rate), when

366

MITCHELL AND NAGEL

critical flow is maintained at the regulating valve. They therefore rejected the claim by Cox et al.144 that the pharmacopeial method is flawed, on the grounds that they drew their conclusion without measuring the flow rate with the DPI in place, assuming that the flow rate entering the system is independent of its resistance (constant mass flow). However, Olsson and Asking145 also noted that, if the compendial method is used unamended, a practical problem arises, since many flow-meters are calibrated for the entering and not the exiting flow that is required to gauge inlet flow rate to the CI correctly. To resolve this concern, they proposed that the relevant flow rate (Qout) simply be calculated from the indicated flow rate (Qin ) and the pressure drop across the flow-meter (DPflow-meter), applying Boyles law

correction for isothermal expansion of an ideal gas: QinPatm Qout 5 } } } (Patm 2 DPflow-meter )

(11)

where Patm is atmospheric pressure. In the case of DPI testing following compendial procedures,1,2 inhaler operation and aerosol formation are intimately linked with the inhalation maneuver simulated by sampling through the CI system, and dead volume is therefore of importance. A large dead volume is normally considered undesirable, as sampling from the DPI only takes place for the required duration to withdraw a fixed volume (4 L) of air from the inhaler mouthpiece. Flow is initiated for the required time by opening a solenoid valve located

FIG. 8. Two methods developed by de Boer et al.147 to vary flow rate rise time when testing DPIs. (Reprinted from Int. J. Pharm. 153, de Boer, A.H., G.K. Bolhuis, D. Gjaltema, and P. Hagerdoorn. Inhalation characteristics and their effects on in vitro drug delivery from dry powder inhalers: Part 3The effect of flow resistance increase rate (FIR) on the in vitro drug release from the Pulmicort 200 Turbuhaler, 6777, Copyright 1997, with permission from Elsevier.)

CASCADE IMPACTORS AND SIZE OF AEROSOLS FROM INHALERS

367

downstream of the CI. The dead volumes associated with the ACI, MSLI and NGI with induction port (contributing 85 mL) amount to approximately 450 mL, 1.2 L, and 1.0 L, respectively. If a pre-separator is used, the additional dead volume of the ACI and NGI increases to approximately 150 mL and 1.0 L, respectively (the pre-separator for the MSLI is stage 1 of this CI). De Boer et al.147 exploited CI system dead volume to study flow rate rise behavior when testing DPIs. They significantly increased the dead volume associated with their MSLI-based system by inserting a 1-L vessel between the MSLI and regulating valve (Fig. 8a). The added volume resulted in a reduced flow rate increase (FIR) up to peak (constant) flow rate when the inhaler was attached and flow initiated through the system. A capillary inserted in the exit tube from the vessel was used to fine tune the flow rate-time profile. Increased FIRs were achieved by creating an under-pressure in the CI, operated without increased dead volume, before connecting and actuating the inhaler (Fig. 8b). The underlying purpose of these modifications was the desire to more closely understand how a high-resistance DPI (Turbuhaler) performs in clinical use, given previous experience with a small group of healthy volunteers.148 Also recognizing the importance of flow rate rise as a function of time when testing DPIs, Chavan and Dalby149 simulated different flow rate ramps that were linear with respect to time, rang-

ing from 100 msec to 3 sec in duration, by regulating the air-flow fed via the DPI using a computer-controlled proportionating valve. Their CI operated at constant flow rate greater than the maximum flow rate achieved via the DPI, following the compendial procedure, with make-up air coming from an inlet that by-passed the inhaler (Fig. 9). Using this approach, they were able to correlate increases in FPF from a Rotahaler DPI with decrease in ramp duration (fastest inhalation), attributing the effect to increased particle de-aggregation and/or the capture of larger aggregates in crevices or regions of low flow within the inhaler at longer ramp durations. Interestingly, they observed an insignificant difference in FPF when 2 L was sampled, rather than the 4-L inhaled volume recommended in the compendial method. Further work is needed, with DPIs having a wide range of resistances, to evaluate this technique more thoroughly. The establishment of the correct CI stage cutpoint sizes at flow rates other than those for which the impactor has been calibrated is also of concern to those testing DPIs by the compendial procedure, as the resistance of the inhaler determines the flow rate at the CI inlet. 1,2 This is appropriate, given the evidence that testing at flow rates appropriate to those achieved by patients is necessary for meaningful comparisons of the performance of this class of inhaler. 146,150 For the ideal inertial separator, it can be shown by application of equations 3 or 4, that the stage cut-

FIG. 9. Method developed by Chavan and Dalby149 for simulating different flow rate ramps for DPI testing by CI. (Reprinted from AAPS PharmSci. 4(2), Chavan, V., and R. Dalby. Novel system to investigate the effects of inhaled volume and rates of rise in simulated inspiratory air flow on fine particle output from a dry powder inhaler, article 6, Copyright 2002, used with permission by the American Association of Pharmaceutical Scientists.)

368

MITCHELL AND NAGEL

point size (d50,1 ) at flow rate (Q1) is related to the cut-point size (d50,ref ) at a reference flow rate (Qref) where calibration data are available, in accordance with: Qref d50,1 5 d50,ref } Q1

1/2

(12)

which is the basis of the calculation provided in the current compendial methods, where d50,ref is fixed at 60 L/min.1,2 This process enables the CI to sample the entire aerosol from the DPI, avoiding concerns about size-related bias, if an alternative such as isokinetic sampling of the aerosol by the CI was to be chosen,93 so that the impactor can be operated only at its design/calibration flow rate. However, equation 12 is only strictly valid when inertial forces dominate the particle separation process so that St is constant.1315 This is not the case when gravity has a significant effect on the size-separation process, as occurs with components of CIs in which particles with dae . 10 mm are being size-separated. 151 Under these circumstances, it may be more appropriate to fit calibration data, if acquired at several flow rates within the range of operation of the CI, by a power law expression of the form: Qref d50,1 5 Y } Q1

(13)

where Qref is a chosen reference condition (generally 60 L/min for DPI testing), and the parameters Y and x are chosen to fit the calibration data. This more general approach was therefore adopted to predict stage cut-point sizes for the NGI.57 Although not based on a model of the underlying physics of impactor operation, the technique is practical, and can also be applied to correct for other non-ideal behavior, such as the effect of the slip correction term describing particle motion, which can be significant with stages that separate particles finer than 0.5 mm. Although most investigators utilize the compendial method for evaluating DPIs, the study by Weuthen et al.152 is of significance, in that it employed an alternative strategy making it possible to characterize inhaler performance at a range of different flow rates from 28.3 to 80 L/min, whilst operating an ACI at a fixed flow rate of 28.3 L/min. In their arrangement, the excess flow was diverted away from the CI via a Y-connector located between the inhaler and induction port, passing through a by-pass line where the volumetric flow rate was measured, to re-join the flow exiting the CI (Fig. 10). A three-way valve located between the CI and vacuum source was used to divert flow from the system so that the inhaler could be attached before making each measurement. Once the valve was switched so that vac-

FIG. 10. CI sampling arrangement for DPI testing developed by Weuthen et al.152 (Reprinted from J. Aerosol Med. 15(3), Weuthen, T., S. Roeder, P. Brand, B. Mllinger, and G. Scheuch. In vitro testing of two formoterol dry powder inhalers at different flow rates, 297303, Copyright 2002, used with permission of Mary Ann Liebert Inc.)

CASCADE IMPACTORS AND SIZE OF AEROSOLS FROM INHALERS

369

uum was applied to the DPI, the entire flow acted on the device within 0.25 sec. In addition to determining the particle size distribution with the CI, the total mass of API emitted from the inhaler could also be quantified by diverting the flow rate downstream of the bypass line to an absolute filter. Although simple in concept, this system appears to lack the advantages of flow stability imposed by the critical orifice, and the particle size-dependent sampling characteristics of the Y-piece interposed between the inhaler and induction port have not been established. Justifying the latter limitation, Weuthen et al.152 noted that particles that impacted in either the Ypiece or induction port are not supposed to reach the lungs, basing the value of their technique on its ability to quantify FPD , 5.8 mm in aerodynamic diameter accurately.

CI CALIBRATION OR STAGE MENSURATION?


In common with other measurement tools concerned with inhaler performance assessments, CI qualification involves a means of providing assurance that performance is to the design specification upon receipt and that it has not degraded with use. Such checks are usually provided by calibration using traceable methods. In the case of a CI, this procedure involves determining each stage collection efficiency-particle size curve to establish its cut size, and this process is most accurately accomplished using monodisperse particles having a range of different sizes in the appropriate range.153 Although unambiguous, calibration with challenge particles of this nature is exceedingly time-consuming, and requires meticulous attention to the methodology to be accurate. It is also dependent upon the accuracy of the calibration particle detecting equipment. Information about internal losses cannot be obtained by methods involving particle counting upstream and downstream of the stage being calibrated, but requires the generation of aerosols comprised of particles containing a chemical marker that can be quantified by appropriate analytical methods.21 For a well-designed CI, the only structure significantly influencing cut size is nozzle diameter, or mean nozzle diameter for a multi-jet stage (equation 4). The U.S. Pharmacopeia, 1 as an alternative to calibration, therefore specifies as part

of system suitability testing for CIs, that the nozzle diameters be measured (stage mensuration). However, no advice is given on the frequency for undertaking this process, other than it should be performed on a regular basis, presumably because the requirement will be formulation and API-recovery method dependent. Nichols,154 in an assessment of standard and modified ACIs for use at high flow rates, recommended that a standardized methodology for stage mensuration be developed for CIs, in association with nozzle specifications that include both mean nozzle diameter and a limit for the range of variation (63 SDs). This approach was taken with the manufacturing specification for the NGI,16 in which a narrow range is specified for nozzle diameters, related to machining tolerances that are specific to each stage.155 Automated optical inspection was proposed for stages whose nozzles are finer than 2.5 mm with the use of go/no-go pingauges as a more precise technique to inspect stages with larger nozzles. These nozzle diameter specifications are associated with the calibration of a representative archival impactor, whose range of nozzle diameters for each stage was purposely chosen to be close to the mid-point of the specification for manufacture. 57 In principle, the user need only confirm after stage mensuration on a regular basis, that the nozzle sizes are still within specification to have confidence in the CI performance, continuing to use the generic archival calibration to define the stage cut sizes. However, recent experience with aluminum ACIs has shown that care is required with the use of go/no go gauges to check stages made from softer metal or where corrosion may have occurred.156 In addition to taking the precaution to limit corrosion by rapid removal of liquid water after cleaning, optical inspection methods may be preferred for these materials, 157 since these examination methods are non-invasive. Although the approach developed for the NGI could in theory be applied to other CIs, until recently there was almost no information available from CI manufacturers to enable standards to be set for stage mensuration. In consequence, users set their own specifications without knowing what effect (if any) deviations outside of these limits might have on stage d50 values, often relying on generic calibration data, such as that supplied by the manufacturer of the ACI. In 1997, Stein and Olson158 highlighted that the nozzle diameters on several stages of 14 aluminum ACIs

370

MITCHELL AND NAGEL

did not correspond to the manufacturers specifications. As a result, cut sizes for individual identical stages were calculated to vary by as much as 0.45 mm. Irregular shaped nozzles formed during manufacture were also identified by microscopy. They then went on to calculate the effect of such variations on those stages where the bulk of the particles would have collected, assuming a model aerosol having MMAD and GSD values of 2.40 mm and 1.70, respectively, was sampled at 28.3 L/min. The mass collected as a percentage of the total sampled mass was estimated to vary from 5.3% to 22.8% for stage 3, 25.3% to 39.2% for stage 4, 29.0% to 38.9% for stage 5, and 5.0% to 10.8% for stage 6. These are substantial variations in the context of inhaler testing, indicating that a more rigorous approach was required for both CI manufacturing specifications and in methods for validating individual CIs. More recently, Stein159 observed from testing undertaken with three different pMDI-produced formulations, that even with ACIs whose nozzles were within the manufacturers specification, large differences in mass of API collected could occur on a stage-by-stage basis. For instance, the mass of API from a solution-based pMDI formulation that collected on stage 6 of ACIs operated at 28.3 L/min ranged from 28.7% to 41.3% of the mass entering the impactor. This finding indicates that it may be impractical to analyze data from this particular CI on an individual stage basis, even when stage mensuration has confirmed that the nozzle dimensions are within specification. In response to these concerns, versions of the ACI manufactured in stainless steel with and without gold-plating have been manufactured in an attempt to improve nozzle manufacturing tolerances as well as to enhance resistance to changes in use caused by wear and/or corrosion. CI manufacturers have also become aware of the need to publish nozzle diameter specifications and some now offer stage mensuration services. Such improvements appear to be necessary, to judge from a recent comparison of aluminum and stainless steel ACIs from two different manufacturers reported by Shelton et al.,160 in which a solution-formulated pMDI was chosen in order to minimize inhaler-based variability. Microscopically visible imperfections in the aluminum impactor nozzles, including non-spherical outline, partial obstruction with debris, ridging of metal at the orifice exit due to the manufacturing process, were associated with small, but signifi-

cantly greater inter-impactor variability in size distribution measurements than equivalent data obtained with CIs manufactured from stainless steel.

CONCLUSION
The CI technique is challenging and requires meticulous attention to detail to make accurate and precise measurements consistently, irrespective of inhaler type. This review has highlighted several aspects of CI design and operation that have an influence on the accuracy of these measurements, but the user has to decide which of these issues should be addressed, based on the application being undertaken. There are two distinct uses for CI-based measurements: (1) the assessment of inhaler quality and (2) the estimation of deposition behavior of the resulting aerosol in the respiratory tract. It would be both an additional burden and divergent from current compendial and regulatory practices to interface a CI with a breathing simulator for the purpose of inhaler batch release testing. Although the ideal result is to establish accurately an absolute measure of the size distribution, a more critical goal in product performance testing is to optimize and maintain consistent measurement capability from batch to batch. Under these circumstances, it is desirable to keep the measurement system as simple as possible. However, the use of the CI in conjunction with a simulator capable of at least reproducing an inhalation maneuver, or better the entire breathing cycle, would be a more appropriate approach to take, if the measurements were being undertaken to understand the likely destination of the inhaler-produced particles in the respiratory tract. A further refinement would be the use of an anatomically correct induction port. Whatever the purpose of the measurement, close attention to details that directly affect CI performancesuch as collection substrate type and coating, leakage, and flow rate controlshould always be a priority to maximize the value of the measurements obtained using this technique.

NOMENCLATURE
C p 5 Cunningham slip correction of particle dp C ae 5 Cunningham slip correction of particle dae C 50 5 Cunningham slip correction factor of a particle of size d50

CASCADE IMPACTORS AND SIZE OF AEROSOLS FROM INHALERS

371

d16 5 size corresponding to the 16th percentile of the particle size distribution or 16th percentile of the stage collection efficiency curve d84 5 size corresponding to the 84th percentile of the particle size distribution or 84th percentile of the stage collection efficiency curve d50 5 stage cut size dac 5 aerodynamic particle diameter dp 5 particle physical diameter D c 5 diameter of the overall cluster of nozzles in a multi-nozzle impactor stage E 5 collection efficiency of an impactor stage E50 5 collection efficiency corresponding to the cut size of an impactor stage FPD 5 fine particle dose per actuation from the inhaler FPF 5 fine particle fraction GSD 5 geometric standard deviation of the particle size distribution GSDstage/pre-sep 5 geometric standard deviation of the stage or pre-separator collection efficiency curve (sharpness-of-cut) n 5 number of nozzles per impactor stage Patm 5 atmospheric pressure DP 5 pressure drop across a flow-meter Q 5 total volumetric flow rate through an impactor stage Q 1 5 flow rate for an impactor measurement Q in 5 volumetric flow rate at the entry to a flowmeter Q out 5 volumetric flow rate at the flow-meter exit Q ref 5 reference flow rate Ref 5 flow Reynolds number S 5 nozzle to collection surface distance St 5 Stokes number St50 5 Stokes number corresponding to d50 T 5 nozzle throat length TED 5 total emitted dose per actuation from the inhaler U 5 average air velocity at nozzle exit of impaction stage W 5 nozzle diameter x 5 adjustable power term to fit impactor calibration data Xc 5 cross-flow parameter Y 5 adjustable coefficient to fit impactor calibration data m 5 air viscosity ra 5 air density rp 5 particle density r0 5 unit density (i.e., 1 g/cm3 ) x 5 dynamic shape factor (1.00 for a spherical particle)

REFERENCES
1. USP 26-NF 21. 2003. Chapter 601physical tests and determinations: aerosols. United States Pharmacopeia, Rockville, MD, 21052123. 2. European Pharmacopeia. 2002. Section 2.9.18 preparations for inhalation: aerodynamic assessment of fine particles. European Pharmacopeia, 3rd ed. Suppl. 2001. Council of Europe, Strasbourg, France, 113124. 3. United States Federal Drug Administration (FDA). 1998. Draft guidance: metered dose inhaler (MDI) and dry powder inhaler (DPI) drug products chemistry, manufacturing and controls documentation, Docket 98D-0997. United States Federal Drug Administration, Rockville, MD. 4. European Agency for the Evaluation of Medicinal Products (EMEA). 2002. Note for guidance on requirements for pharmaceutical documentation for pressurised metered dose inhalation products (docket CPMP/QWP/2845/00). European Agency for the Evaluation of Medicinal Products, London, UK. 5. Hinds, W.C. 1999. Properties, Behavior, and Measurement of Airborne Particles, 2nd ed. Wiley-Interscience, New York. 6. Rudolph, G., R. Kobrich, and W. Stahlhofen. 1990. Modeling and algebraic formulation of regional aerosol deposition in man. J. Aerosol Sci. 21:306406. 7. Heyder, J., and M.U. Svartengren. 2002. Basic principles of particle behavior in the human respiratory tract. In H. Bisgaard, C. OCallaghan, and G.C. Smaldone eds. Drug Delivery to the Lung. Marcel Dekker, New York, 2145. 8. Hering, S.V., R.C. Flagan, and S.K. Friedlander. 1978. Design and evaluation of a new low-pressure impactor. Environ. Sci. Technol. 12:667673. 9. Marple, V.A., K.L. Rubow, and B.A. Olson. 2001. Inertial, gravitational, centrifugal, and thermal collection techniques. In P.A. Baron and K.Willeke, eds. Aerosol Measurement: Principles, Techniques and Applications, 2nd ed. Wiley Interscience, New York, 229260. 10. Suman, J.D., B.L. Laube, and R. Dalby. 1999. Comparison of nasal deposition and clearance of aerosol generated by a nebulizer and an aqueous spray pump. Pharm. Res. 16:16481652. 11. Suman, J.D., B.L. Laube, and R. Dalby. 2002. Documenting nasal bioequivalence from in vitro characteristics to physiologic response. In R.N. Dalby, P.R. Byron, J. Peart, et al., eds. Respiratory Drug Delivery VIII. Davis Horwood International, Raleigh, NC, 691693. 12. Tillery, M., and R. Buchan. 2002. Determination of large aerosol particle size by elutriation. Appl. Occup. Environ. Hyg. 17:717722. 13. Marple, V.A., and B.Y.H. Liu. 1974. Characteristics of laminar jet impactors. Environ. Sci. Technol. 8: 648654. 14. Marple, V.A., and K. Willeke. 1976. Inertial im-

372
pactors: theory, design and use. In B.Y.H. Liu, ed. Fine Particles. Academic Press, New York, 411466. Rader, D.J., and V.A. Marple. 1985. Effect of ultraStokesian drag and particle interception on impactor characteristics. Aerosol Sci. Technol. 4:141156. Marple, V.A., D.L. Roberts, F.J. Romay, et al. 2003. Next generation pharmaceutical impactor. Part 1: Design. J. Aerosol Med. 16:283299. OShaughnessy, P.T., and O.G. Raabe. 2003. A comparison of cascade impactor data reduction methods. Aerosol Sci. Technol. 37:187200. Marple, V.A., D.L. Roberts, and F.J. Romay. 2000. Design of the next generation pharmaceutical impactor. In Drug Delivery to the LungsXI. The Aerosol Society, London, UK, 127130. Rader, D.J., and V.A. Marple. 1984. Effect of gravitational forces on the calculation of impactor efficiency curves. In B.Y.H. Liu, D.Y.H. Pui, and H.J. Fissan, eds. Aerosols. Elsevier, New York, 123126. Marple, V.A. 2002. Private communication, University of Minnesota, MN. Marple, V.A., B.A. Olson, and N.C. Miller. 1998. The role of inertial particle collectors in evaluating pharmaceutical aerosol systems. J. Aerosol Med. 11: S139153. Fang, C.P., V.A. Marple, and K.L. Rubow. 1991. Influence of cross-flow on particle collection characteristics of multi-nozzle impactors. J. Aerosol Sci. 22: 403415. Berg, E., J.O. Svensson, and L. Asking. 2002. MMAD based on dose to impactor rather than on delivered dose. In R.N. Dalby, P.R. Byron, Peart, J., et al., eds. Respiratory Drug Delivery VIII. Davis Horwood International, Raleigh, NC, 339342. Canadian Standards Association. 2003. Spacers and holding chambers for use with metered-dose inhalers. CAN/CSA Z264.1-02. Canadian Standards Association, Toronto, Canada. Stein, S.W., B.J. Gabrio, D. Oberreit, et al. 2002. An evaluation of mass-weighted size distribution measurements with the model 3320 aerodynamic particle sizer. Aerosol Sci. Technol. 36:845854. Dolovich, M. 1991. Measurement of particle size characteristics of metered dose inhaler (MDI) aerosols. J. Aerosol Med. 4:251263. Malton, C.A., G.W. Hallworth, and J.M. Padfield. 1982. The association and particle size distribution of drug and surfactant discharged from a metereddose inhalation aerosol. J. Pharm. Pharmacol. 34:65P. Dunbar, C.A., and A.J. Hickey. 2000. Evaluation of probability density functions to approximate particle size distributions of representative pharmaceutical inhalers. J. Aerosol Sci. 31:813831. Thiel, C.G. 1998. Can in vitro particle size measurements be used to predict pulmonary deposition of aerosol from inhalers? J. Aerosol Med. 11:4352. Product Quality Research Institute (PQRI). 2002. Work plan: investigation of an optimized chisquared method for comparing particle size distrib-

MITCHELL AND NAGEL


ution profiles obtained by cascade impactors with specific reference to equivalence testing or orally inhaled and pressurized nasal drug products. [On-line] Available: www.pqri.org/minutes/pdfs/ dptc/psdpcwg/workplan02.pdf. Clark, A.R., and N. Kadrichu. 2000. Comparing pharmaceutical aerosol particle size distributions. In R.N. Dalby, P.R. Byron, Farr, S.J., et al., eds. Respiratory Drug Delivery VII. Serentec Press, Raleigh, NC, 181189. Olsson, B., J. Aiache, H. Bull, et al. 1996. The use of inertial impactors to measure the fine particle dose generated by inhalers. Pharmeuropa 8:291298. Cripps, A., M. Riebe, M. Schulze, et al. 2000. Pharmaceutical transition to non-CFC pressurized metered dose inhalers. Respir. Med. 94:39. Newhouse, M.T. 1998. The current laboratory determination of respirable mass is not clinically relevant. J. Aerosol Med. 11:122132. Federal Drug Administration. 2002. Guidance for industry: nasal spray and inhalation solution, suspension, and spray drug productschemistry, manufacturing, and controls documentation. Docket 99D-1454. Center for Drug Evaluation and Research, Rockville, MD. Poochikian, G., and C.M. Bertha. 2002. Regulatory view on current issues pertaining to inhalation drug products. In R.N. Dalby, P.R. Byron, Peart, J., et al., eds. Respiratory Drug Delivery VIII. Davis Horwood International, Raleigh, NC, 159164. Chemistry Materials and Controls Specifications Technical Team. 2000. Initial assessment of the ITFG/IPAC aerodynamic particle size distribution database by the CMC Specifications Technical Team of the ITFG/IPAC collaboration. International Pharmaceutical Aerosol Consortium on Regulation and Science (IPAC-RS) [On-line]. Available: www. ipacrs.com/particle_size.html. Product Quality Research Institute (PQRI). 2002. Work plan: establishment of the appropriate use of the particle size distribution mass balance determined by cascade impactor for orally inhaled and nasal drug products. [On-line]. Available: www. pqri.org/minutes/pdfs/dptc/psdmbwg/workplan02.pdf. Christopher, D., P. Curry, B. Doub, K., et al. 2003. Considerations for the development and practice of cascade impaction testing including a mass balance failure investigation tree. J. Aerosol Med. 16:235247. Ahrens, R., C. Lux, T. Bahl, et al. 1995. Choosing the metered-dose inhaler spacer or holding chamber that matches the patients need: evidence that the specific drug being delivered is an important consideration. J. Allergy Clin. Immunol. 96:288294. Mitchell, J.P., M.W. Nagel, and J.L. Rau. 1999. Performance of large volume versus small volume holding chambers with chlorofluorocarbon-albuterol and hydrofluoroalkane-albuterol sulfate. Respir. Care 44:3844.

15.

16.

31.

17.

32.

18.

33.

19.

34.

20. 21.

35.

22.

36.

23.

37.

24.

25.

38.

26.

27.

39.

28.

40.

29.

41.

30.

CASCADE IMPACTORS AND SIZE OF AEROSOLS FROM INHALERS


42. Mitchell, J.P., and M.W. Nagel. 2000. Spacer and holding chamber testing in vitro: a critical analysis with examples. In R.N. Dalby, P.R. Byron, S.J. Farr, et al., eds. Respiratory Drug Delivery VII. Serentec Press, Raleigh, NC, 265273. 43. Olsson, B., L. Asking, and M. Johansson. 1998. Choosing a cascade impactor. In R.N. Dalby, P.R. Byron, S.J. Farr, et al., eds. Respiratory Drug Delivery VI. Interpharm Press, Buffalo Grove, IL, 133138. 44. Subba Rao, G.N., D.W. Banick, and D.M. Pacenti. 1997. Particle size distribution of a suspension aerosol using Andersen and Marple-Miller cascade impactors. Pharm. Res. 14:12721274. 45. Andersen, A. 1958. A sampler for respiratory health assessment. Am. Ind. Hyg. Assoc. J. 27:160165. 46. Nichols, S.C., D.R. Brown, and M. Smurthwaite. 1998. New concept for the variable flow rate Andersen cascade impactor and calibration data. J. Aerosol Med. 11:133138. 47. Nichols, S.C. 2000. Calibration and mensuration issues for the standard and modified impactor. Pharmeuropa 12:585. 48. Mitchell, J.P., P.A. Costa, and S. Waters. 1987. An assessment of an Andersen Mark-II cascade impactor. J. Aerosol Sci. 19:213221. 49. Rao, A.K., and K.T. Whitby. 1978. Non-ideal collection characteristics of inertial impactorssingle stage impactors and solid particles. J. Aerosol Sci. 9:7786. 50. Rao, A.K., and K.T. Whitby. 1978. Non-ideal collection characteristics of inertial impactorscascade impactors. J. Aerosol Sci. 9:87100 51. Miller, N.C. 1994. A cascade impactor for aerodynamic size measurement for MDIs and powder inhalers. In P.R. Byron, R.N. Dalby, and S.J. Farr, eds. Respiratory Drug Delivery IV. Interpharm Press, Buffalo Grove, IL, 342343. 52. Marple, V.A., B.A. Olson, and N.C. Miller. 1995. A low-loss cascade impactor with stage collection cups: calibration and pharmaceutical inhaler applications. Aerosol Sci. Technol. 22:124134. 53. Olson, B.A., V.A. Marple, J.P. Mitchell, et al. Development and calibration of a low-flow version of the Marple-Miller impactor. Aerosol Sci. Technol. 29: 307314. 54. Hindle, M., P.R. Byron, and N.C. Miller. 1996. Cascade impaction methods for dry powder inhalers using the high flow rate Marple-Miller impactor. Int. J. Pharm. 134:137146. 55. Vaughan, N.P. 1989. The Andersen impactor: calibration, wall losses and numerical simulation. J. Aerosol Sci. 20:6790. 56. LeBelle, M.J. S.J. Graham, R.K. Pike, et al. 1997. Metered-dose inhalers. II. Particle size measurement variation. Int. J. Pharm. 151:209221. 57. Marple, V.A., B.A. Olson, K. Santhanakrishnan, et al. 2003. Next generation pharmaceutical impactor. Part II. Calibration. J. Aerosol Med. 16:301324.

373

58. May, K.R. 1966. Multi-stage liquid impinger. Bact. Rev. 30:559570. 59. Bell, J.H., K. Brown, and J. Glasby. 1973. Variation in delivery of isoprenaline from various pressurized inhalers. J. Pharm. Pharmacol. 25:32P36P. 60. Asking, L., and B. Olsson. 1997. Calibration at different flow rates of a multistage liquid impinger. Aerosol Sci. Technol. 27:3949. 61. Yu, C.D., R.E. Jones, and M. Henesian. 1984. Cascade impactor method for the droplet size characterization of a metered-dose nasal spray. J. Pharm. Sci. 73:344348. 62. Mitchell, R.I., and J.M. Pilcher. 1958. Design and calibration of an improved cascade impactor for size analysis of aerosols. Presented at the 5th Air Cleaning Conference, Atomic Energy Commission, Washington, D.C. 63. Tzou, T.Z. 1999. Aerodynamic particle size of metered dose inhalers determined by the quartz crystal microbalance and the Andersen impactor. Int. J. Pharm. 186:7179. 64. Tzou, T.Z., and J.M. Elvecrog. 1995. Comparing the aerodynamic particle size of MDIs measured by the quartz crystal microbalance cascade impactor and the Andersen cascade impactor. Pharm. Res. 12:S181. 65. Horton, K.D., M.H.E. Ball, and J.P. Mitchell. 1992. The calibration of a California Measurements PC-2 quartz crystal cascade impactor. J. Aerosol Sci. 23:505524. 66. Peart, J., Magyar, C and P.R. Byron. 1998. Aerosol electrostaticsmetered dose inhalers (MDIs): reformulation and device design issues. In R.N Dalby, P.R. Byron, and S.J. Farr, eds. Respiratory Drug Delivery VI. Interpharm Press, Buffalo Grove, IL, 227233. 67. Byron, P.R., J. Peart, and J.N. Staniforth. 1997. Aerosol electrostatics. I. Properties of fine powders before and after aerosolization by dry powder inhalers. Pharm. Res. 14:698705. 68. Chow, H.Y., and T.T. Mercer. 1971. Charges on droplets produced by atomization of solutions. Am. Ind. Hyg. Assoc. J. 31:247255. 69. Dunbar, C.A., and A.J. Hickey. 1999. Selected parameters affecting characterization of nebulized aqueous solutions by inertial impaction and comparison with phase-Doppler analysis. Eur. J. Pharm. Biopharm. 48:171177. 70. Dolovich, M.B. 2002. Assessing nebulizer performance. Respir. Care 47:12901304. 71. Clark, A.R. 1995. The use of laser diffraction for the evaluation of the aerosol clouds generated by medical nebulizers. Int. J. Pharm. 115:6978. 72. Dahlbck, M. 1994. Behavior of nebulizing solutions and suspensions. J. Aerosol Med. 7:1317. 73. United States Federal Drug Administration (FDA). 1993. Reviewer guidance for nebulizers, metered dose inhalers, spacers and actuators. United States Federal Drug Administration (CDRH), Rockwille, MD. 74. Dennis, J.H., Pieron, C.A., and O. Nerbrink. 2000. Standards in assessing in vitro nebulizer performance. Eur. Respir. Rev. 10:178182.

374
75. Finlay, W.H., and K.W. Stapleton. 1999. Undersizing of droplets from a vented nebulizer caused by aerosol heating during transit through an Andersen impactor. J. Aerosol Sci. 30:105109. 76. Jauernig, J., M. Hug, M. Knoch, et al. 2002. Contribution to the aerodynamic particle size assessment of nebulized drugs using the next generation impactor (NGI). In Drug Delivery to the Lungs XIII. Aerosol Society, London, UK, 4447. 77. Rubow, K.L., V.A. Marple, J. Olin, et al. 1987. A personal cascade impactor: design, evaluation and calibration. Am. Ind. Hyg. Assoc. J. 48:532538. 78. European Committee for Standardization. 2001. Respiratory therapy equipmentPart 1: Nebulizing systems and their components. EN 135441. CEN, Brussels, Belgium. 79. Dennis, J.H. 2002. Standardization issues: in vitro assessment of nebulizer performance. Respir. Care 47:14451458. 80. Boe, J., J.H. Dennis, and B.R. ODriscoll. 2001. European Respiratory Society Guidelines on the use of nebulizers. Eur. Respir. J. 18:228242. 81. Smith, M. 2000. Ictus automated impactor particle sizing. In R.N. Dalby, P.R. Byron, S.J. Farr, et al., eds. Respiratory Drug Delivery VII. Serentec Press, Raleigh, NC, 451454. 82. Smith, M.P. 2001. Automation of the next generation impactorfirst the Andersen, now the NGI. In Drug Delivery to the Lungs XI. Aerosol Society, London, UK, 131134. 83. Roberts, D.L. 2003. Private communication, MSP Corporation, Minneapolis, MN. 84. Miller, N.C., D.L. Roberts, and V.A. Marple. 2002. The service head approach to automating the next generation pharmaceutical impactor: proof of concept. In R.N. Dalby, P.R. Byron, J. Peart, et al., eds. Respiratory Drug Delivery VIII. Davis Horwood International, Raleigh, NC, 521523. 85. Fransson, K., M. Persson, and M. Svensson. 2002. Sample preparation tools for the next generation pharmaceutical impactor. In Drug Delivery to the Lungs XIII. Aerosol Society, London, UK, 127130. 86. Dolovich, M., and R. Rhem. 1998. Impact of oropharyngeal deposition on inhaled dose. J. Aerosol Med. 11:112115. 87. Hickey, A.J. and R.M. Evans. 1996. Aerosol generation for propellant-driven metered dose inhalers. In A.J. Hickey, ed. Inhalation Aerosols: Physical and Biological Basis for Therapy. Marcel Dekker New York, 417439. 88. Stein, S.W., and B.J. Gabrio. 2000. Understanding throat deposition during cascade impactor testing. In R.N. Dalby, P.R. Byron, Farr, S.J., et al., eds. Respiratory Drug Delivery VII. Serentec Press, Raleigh, NC, 573576. 89. Van Oort, M., and B. Downey. 1996. Cascade impaction of MDIs and DPIs: Induction port, inlet cone, and pre-separator lid designs recommended for in-

MITCHELL AND NAGEL


clusion in the general test chapter Aerosols k601l. Pharm. Forum. 22:22042210. Van Oort, M., and K. Truman. What is respirable dose? J. Aerosol Med. 11S1:8996. Van Oort, M., R. Gollmar, and R. Bohinski. 1994. Effects of sampling chamber volume and geometry on aerodynamic size distributions of metered-dose inhalation aerosols measured with the Andersen cascade impactor. Pharm. Res. 11:604. Sequeira, J., J. Berry, S. Sharpe, et al. 2002. A comparison of metered dose inhaler particle size distribution by Andersen cascade impaction using two types of entry ports. In R.N. Dalby, P.R. Byron, J. Peart, et al., eds. Respiratory Drug Delivery VIII. Davis Horwood International, Raleigh, NC, 573576. Williams, J.A., and C.L. Witham. 1994. A new cascade impactor induction port. In P.R. Byron, R.N. Dalby, and S.J. Farr, eds. Respiratory Drug Delivery IV. Interpharm Press, Buffalo Grove, IL, 348350. Byron, P.R. 1994. Compendial dry powder testing: USP perspectives. In P.R. Byron, R.N. Dalby, and S.J. Farr, eds. Respiratory Drug Delivery IV. Interpharm Press, Buffalo Grove, IL, 153162. Van Oort, M., B. Downey, and W. Roberts. 1996. Verification of operating the Andersen cascade impactor at different flow rates. Pharm. Forum 22:22112215. Brouet, G., S. Burel, and J.-C. Gilles. 2002. A modified Andersen cascade impactor throat for testing breath actuated inhalers. In R.N. Dalby, P.R. Byron, J. Peart, et al., eds. Respiratory Drug Delivery VIII. Davis Horwood International, Raleigh, NC, 747750. Borgstrm, L. 1999. In vitro, ex vivo, in vivo veritas. J. Allergy 54S49:8892. Massoud, O., G.P. Martin, C. Marriott, et al. 2002. The in vitro assessment of aerosolized drug deposition using novel oropharyngeal models. In Drug Delivery to the Lungs XIII. Aerosol Society, London, UK, 2730. Berg, E. 1995. In vitro properties of pressurized metered dose inhalers with and without spacer devices. J. Aerosol Med. 8S3:311. Olsson, B., L. Borgstrm, L. Asking, et al. 1996. Effect of inlet throat on the correlation between measured fine particle dose and lung deposition. In R.N Dalby, P.R. Byron, and S.J. Farr, eds. Respiratory Drug Delivery V. Interpharm Press, Buffalo Grove, IL, 273281. Berg, E., C. Rossberg, E. Lindqvist, et al. 1998. Dose exiting models of the human throata comparison of five different models. In Drug Delivery to the Lungs IX. Aerosol Society, London, UK, 2932. Srichana, T., G.P. Martin, and C. Marriott. 2000. A human oral-throat cast integrated with a twin-stage impinger for evaluation of dry powder inhalers. J. Pharm. Pharmacol. 52:771778. Smaldone, G.C., and P.N. LeSouef. 2002. Nebulization: the device and clinical considerations. In H. Bisgaard, C. OCallaghan and G.C. Smaldone, eds. Drug Delivery to the Lung. Marcel Dekker, New York, 269302.

90. 91.

92.

93.

94.

95.

96.

97. 98.

99.

100.

101.

102.

103.

CASCADE IMPACTORS AND SIZE OF AEROSOLS FROM INHALERS


104. Sangwan, S., R. Condos, and G.C. Smaldone. 2003. Lung deposition and respirable mass during wet nebulization. J. Aerosol Med. (submitted). 105. Miller, N.C., and A.M. Purrington. 1996. A cascade impactor entry port for MDI sprays with collection characteristics imitating a physical model of the human throat. Pharm. Res. 13:391397. 106. Mitchell, J.P., M.W. Nagel, C.C. Doyle, et al. 2002. The effect of humidification on the size distribution of metered dose inhaler aerosols to the mechanically ventilated patient. In Drug Delivery to the Lungs XIII. Aerosol Society, London, UK, 3639. 107. Sethuraman, V.V., and A.J. Hickey. 2001. Evaluation of pre-separator performance for the 8-stage nonviable Andersen impactor. AAPS PharmSciTech. [Online]. Available: http://www.aapspharmscitech.org/. 108. Roberts, D.L., F.A. Romay, V.A. Marple, et al. 2000. A high-capacity pre-separator for cascade impactors. In R.N. Dalby, P.R. Byron, Farr, S.J., et al. eds. Respiratory Drug Delivery VII. Serentec Press, Raleigh, NC, 443445. 109. Fink, J.B., and R. Dhand. 2001. Laboratory evaluation of metered-dose inhalers with models that simulate interaction with the patient. Respir. Care Clin. North Am. 7:303317. 110. Finlay, W.H. 1998. Inertial sizing of aerosol inhaled during pediatric tidal breathing from an MDI with attached holding chamber. Int. J. Pharm. 168:147152. 111. Finlay, W.H., and P. Zuberbuhler. 1999. In vitro comparison of salbutamol hydrofluoroalkane (Airomir) metered dose inhaler aerosols inhaled during pediatric tidal breathing from five valved holding chambers. J. Aerosol Med. 12:285291. 112. Finlay, W.H., and P. Zuberbuhler. 1999. In vitro comparison of beclomethasone and salbutamol metereddose inhaler aerosols inhaled during pediatric tidal breathing from four valved holding chambers. Chest 114:16761680. 113. Foss, S.A., and J.W. Keppel. 1999. In vitro testing of MDI spacers: a technique for measuring respirable dose output with actuation in-phase or out-of-phase with inhalation. Respir. Care. 44:14741485. 114. Mitchell, J.P., and M.W. Nagel. 1997. In vitro performance testing of three small volume holding chambers under conditions that correspond with use by infants and small children. J. Aerosol Med. 10:341349. 115. Brindley, A., R.M. Marriott, B.S. Sumby, et al. 1994. The electronic lung: a novel tool for the characterization of inhalation devices. J. Pharm. Pharmacol. 45:135. 116. Burnell, P.K.P., A. Malton, K. Reavill, et al. 1998. Design, validation and initial testing of the electronic lung device. J. Aerosol Sci. 29:10111025. 117. Finlay, W.H., and M.G. Gehmlich. 2000. Inertial sizing of aerosol inhaled from two dry powder inhalers with realistic breath patterns versus constant flow rates. Int. J. Pharm. 210:8395. 118. Mitchell, J.P., M.W. Nagel, K.J. Wiersema, et al. 2000.

375

119.

120.

121. 122.

123.

124.

125.

126.

127.

128.

129.

130.

Performance of large and small volume valved holding chambers as a function of flow rate. In Drug Delivery to the Lungs XI. Aerosol Society, London, UK, 5255. Smith, K.J., H.-K. Chan, and K.F. Brown. 1998. Influence of flow rate on aerosol particle size distributions from pressurized and breath-actuated inhalers. J. Aerosol Med. 11:231245. Federal Drug Administration. 1999. Draft guidance for industry: bioavailability and bioequivalence studies for nasal aerosols and nasal sprays for local action. Document 2070. Center for Drug Evaluation and Research, Rockville, MD. Ranucci, J. 1992. Dynamic plume-particle size analysis using laser diffraction. Pharm. Technol. 16:108114. Eck, C.R., T.F. McGrath, and A.G. Perlwitz. 2000. Droplet size distributions in a solution nasal spray. In R.N. Dalby, P.R. Byron, Farr, S.J., et al., eds. Respiratory Drug Delivery VII. Serentec Press, Raleigh, NC, 475478. Adams, W.P., G.J.P. Singh, and R.L. Williams. 1998. Nasal inhalation aerosols and metered dose spray pumps: FDA bioequivalence issues. In R.N. Dalby, P.R. Byron, and S.J. Farr, eds. Respiratory Drug Delivery VI. Interpharm Press, Buffalo Grove, IL, 219225. Chien, Y.W. 1992. Nasal drug delivery and delivery systems. In Y.W. Chien, ed. Novel Drug Delivery Systems, 2nd ed. Marcel Dekker, New York, 229268. Harrison, L.I. 2000. Commentary on the FDA draft guidance for bioequivalence studies for nasal aerosols and nasal sprays: an industry view. J. Clin. Pharmacol. 40:701707. Guo, Y., J.D. Suman, and R. Dalby. 2002. Influence of entry port dimensions and air flow rates on aqueous nasal spray deposition in the Andersen cascade impactor. Presented at the Annual Meeting of American Association of Pharmaceutical Scientists, Toronto, Canada. Suman, J.D. 2002. Validity of in vitro tests for nasal delivery systems as surrogates for in vivo deposition pattern and biologic response. Ph.D. dissertation. University of Maryland, College Park, MD. Doub, W.H., and W.P. Adams. 2002. Measurement of drug in fine particles from aqueous nasal spray by cascade impactor. Presented at the Annual Meeting of American Association of Pharmaceutical Scientists, Toronto, Canada. Ostrander, K.D., R.L. Mueller, J.R. Swanson, et al. 1995. A novel induction port for use with impaction devices to assess the aerodynamic size distribution of nasally administered medicants. Pharm. Technol. 19:98106. Leung, S., D. Velasquez, and D. Schultz. 1997. An estimation of deposition for nasal metered dose inhalers with in vitro models. Presented at the Annual Meeting of American Association of Pharmaceutical Scientists. Available: www.3m.com/us/healthcare/ manufacturers/dds/pdf/an_estimate.pdf.

376
131. Mitchell, J.P. 2000. The next generation impactor (NGI): results from the evaluation of prototype instruments with pressurized metered dose inhaler (pMDI)based formulations. In Drug Delivery to the Lungs XI. Aerosol Society, London, UK, 223226. 132. Shrubb, I. 2000. The next generation impactor (NGI): results from the evaluation of prototype instruments with dry powder inhaler (DPI)based formulations. In Drug Delivery to the Lungs XI. Aerosol Society, London, UK, 227230. 133. Esmen, N.A., and T.C. Lee. 1980. Distortion of cascade impactor measured size distribution due to bounce and blow-off. Am. Ind. Hyg. Assoc. J. 41: 410419. 134. Dunbar, C.A., A.J. Hickey, and P. Holzner. 1998. Dispersion and characterization of pharmaceutical dry powder aerosols. KONA 16:745. 135. Podczeck, F. 1997. Optimization of the operation conditions of an Andersen cascade impactor and the relationship to centrifugal adhesion measurements to aid in the development of dry powder inhalations. Int. J. Pharm. 149:5161. 136. Graham, S.J., R.C. Lawrence, E.D. Ormsby, et al. 1995. Particle size distribution of single and multiple sprays of salbutamol metered-dose inhalers (MDIs). Pharm. Res. 12:13801384. 137. Nasr, M.M., D.L. Ross, and N.C. Miller. 1997. Effect of drug load and plate coating on the particle size distribution of a commercial albuterol metered dose inhaler (MDI) determined using the Andersen and Marple-Miller impactors. Pharm. Res. 14:14371443. 138. Nasr, M.M., and J.F. Allgire. 1995. Loading effect on particle size measurements by inertial sampling of albuterol metered dose inhalers. Pharm. Res. 12: 16771681. 139. El-Araud, K.A., B.J. Clark, C. Kaahwa, et al. 1998. The effect of dose on the characterization of aerodynamic particle size distributions of beclomethasone dipropionate metered-dose inhalers. J. Pharm. Pharmacol. 50:10811085. 140. Young, J.Y. 1995. Particle Sampling Using Cascade Impactors: Some Practical Application Issues. American Industrial Hygiene Association, Fairfax, VA. 141. Chen, B.T., and W. John. 2001. Instrument calibration. In P.A. Baron and K. Willeke, eds. Aerosol Measurement: Principles, Techniques and Applications, 2nd ed. Wiley Interscience, New York, 627666. 142. Barnes, A.R., and S. Nash. 1997. Beclomethasone dipropionate 250 mg per dose metered dose inhalers: effect of volumatic spacer on potentially respirable dose. Int. J. Pharm. 157:145152. 143. Peart, J., P.R. Byron, T.S. Staehler, et al. 1997. Pressure-drop measurements made during testing of dry powder inhalers. Pharm. Forum 23:35433546. 144. Cox, R.L., H.K. Versteeg, and M.J. Shott. 2001. Measurement and setting of flow rates in pharmaceuti-

MITCHELL AND NAGEL


cal aerosol dispersion testing. In Drug Delivery to the Lungs XII. Aerosol Society, London, UK, 143146. Olsson, B., and L. Asking. 2002. Methods of setting and measuring flow rates in pharmaceutical impactor experiments. In Drug Delivery to the Lungs XIII. Aerosol Society, London, UK, 168171. Clark, A.R., and A.M. Hollingworth. 1993. The relationship between powder inhaler resistance and peak inspiratory conditions in healthy volunteers implications for in vitro testing. J. Aerosol Med. 6: 99110. de Boer, A.H., G.K. Bolhuis, D. Gjaltema, et al. 1997. Inhalation characteristics and their effects on in vitro drug delivery from dry powder inhalers: Part 3 The effect of flow resistance increase rate (FIR) on the in vitro drug release from the Pulmicort 200 Turbuhaler. Int. J. Pharm. 153:6777. de Boer, J.H., H.M.I. Winte, and C.F. Lerk. 1995. Inhalation characteristics and air flow resistance of dry powder inhalers as preferred by healthy subjects. Eur. Respir. J. 8S19:425. Chavan, V., and R. Dalby. 2002. Novel system to investigate the effects of inhaled volume and rates of rise in simulated inspiratory air flow on fine particle output from a dry powder inhaler. AAPS Pharm/www.aapspharmSci., [On-line]. Available: http:/ sci.org/. Clark, A.R. 1995. Medical aerosol inhalers: past, present and future. Aerosol Sci. Technol. 22:374391. Huang, C.-H. and C.-J. Tsai. 2001. Effect of gravity on particle collection efficiency of inertial impactors. J. Aerosol Sci. 32:375387. Weuthen, T., S. Roeder, P. Brand, et al. 2002. In vitro testing of two formoterol dry powder inhalers at different flow rates. J. Aerosol Med. 15:297303. Mitchell, J.P. 1998. Aerosol generation and instrument calibration. In I. Colbeck, ed. Physical and Chemical Properties of Aerosols. Blackie Academic and Professional, London, UK, 3179. Nichols, S.C. 2000. Calibration and mensuration issues for the standard and modified Andersen cascade impactor. Pharm. Forum 26:14661469. Roberts, D.L., F.J. Romay, and V.A. Marple. 2001. Nozzle examination methods for the next generation pharmaceutical impactor. In Drug Delivery to the Lungs XII. Aerosol Society, London, UK, 6669. Svensson, M., G. Pettersson, and L. Asking. 2002. Stage mensuration of the Andersen impactorpitfalls and recommendations. In Drug Delivery to the Lungs XIII. Aerosol Society, London, UK, 188191. Carter, I., F. Chambers, and A. Walsh. 2002. Use of the Mitutoyo Quick Vision 404 system to mensurate the critical dimensions of impactors. In Drug Delivery to the Lungs XIII. Aerosol Society, London, UK, 229232. Stein, S.W., and B.A. Olson. 1997. Variability in size

145.

146.

147.

148.

149.

150. 151.

152.

153.

154.

155.

156.

157.

158.

CASCADE IMPACTORS AND SIZE OF AEROSOLS FROM INHALERS


distribution measurements obtained using multiple Andersen mark II cascade impactors. Pharm. Res. 14:17181725. 159. Stein, S.W. 1999. Size distribution measurements of metered dose inhalers using Andersen mark II cascade impactors. Int. J. Pharm. 186:4352. 160. Shelton, C., B. Woodrow, W. Holberg, et al. 2002. Performance comparison of Copley and Andersen 8-stage cascade impactors. In Drug Delivery to the Lungs XIII. Aerosol Society, London, UK, 176179.

377

Received on January 31, 2003; in final form, March 27, 2003 Reviewed by: Leon Gradon, Ph.D. Address reprint requests to: Jolyon P. Mitchell, Ph.D. Trudell Medical International 725 Third St. London, Ontario, Canada N5V 5G4 E-mail: jmitchell@trudellmed.com

You might also like