throbber
AAPS PharmSciTech 2003; 4 (4) Article 54 (http://www.aapspharmscitech.org).
`Aerodynamic Particle Size Analysis of Aerosols from Pressurized Metered-
`Dose Inhalers: Comparison of Andersen 8-Stage Cascade Impactor, Next
`Generation Pharmaceutical Impactor, and Model 3321 Aerodynamic Par-
`ticle Sizer Aerosol Spectrometer
`Submitted: July 14, 2003; Accepted: September 2, 2003
`Jolyon P. Mitchell,1 Mark W. Nagel,1 Kimberly J. Wiersema,1 and Cathy C. Doyle1
`1Trudell Medical International, 725 Third Street, London ON, Canada N5V 5G4
`
`
`ABSTRACT
`The purpose of this research was to compare three dif-
`ferent methods for the aerodynamic assessment of (1)
`chloroflurocarbon
`(CFC)
`-fluticasone propionate
`(Flovent), (2) CFC-sodium cromoglycate (Intal), and
`(3)
`hydrofluoroalkane
`(HFA)
`-beclomethasone
`dipropionate (Qvar) delivered by pressurized metered
`dose inhaler. Particle size distributions were compared
`determining mass median aerodynamic diameter
`(MMAD), geometric standard deviation (GSD), and
`fine particle fraction <4.7 µm aerodynamic diameter
`(FPF<4.7 µm). Next Generation Pharmaceutical Impactor
`(NGI)-size distributions for Flovent comprised finer
`particles than determined by Andersen 8-stage impac-
`tor (ACI) (MMAD = 2.0 ± 0.05 µm [NGI]; 2.8 ± 0.07
`µm [ACI]); however, FPF<4.7 µm by both impactors was
`in the narrow range 88% to 93%. Size distribution
`agreement for Intal was better (MMAD = 4.3 ± 0.19
`µm (NGI), 4.2 ± 0.13 µm (ACI), with FPF<4.7 µm rang-
`ing from 52% to 60%. The Aerodynamic Particle Sizer
`(APS) undersized aerosols produced with either formu-
`lation (MMAD = 1.8 ± 0.07 µm and 3.2 ± 0.02 µm for
`Flovent and Intal, respectively), but values of FPF<4.7 µm
`from the single-stage impactor (SSI) located at the inlet
`to the APS (82.9% ± 2.1% [Flovent], 46.4% ± 2.4%
`[Intal]) were fairly close to corresponding data from the
`multi-stage impactors. APS-measured size distributions
`for Qvar (MMAD = 1.0 ± 0.03 µm; FPF<4.7 µm = 96.4%
`± 2.5%), were in fair agreement with both NGI
`(MMAD = 0.9 ± 0.03 µm; FPF<4.7 µm = 96.7% ± 0.7%),
`and ACI (MMAD = 1.2 ± 0.02 µm, FPF<4.7 µm = 98% ±
`0.5%), but FPF<4.7 µm from the SSI (67.1% ± 4.1%) was
`
`Corresponding Author: Jolyon P. Mitchell, Trudell
`Medical International, 725 Third Street, London ON,
`Canada N5V 5G4. Tel: (519) 455-7060 ext.2206; Fax:
`(519) 455-9053;Email: jmitchell@trudellmed.com
`
`
`lower than expected, based on equivalent data obtained
`by the other techniques. Particle bounce, incomplete
`evaporation of volatile constituents and the presence of
`surfactant particles are factors that may be responsible
`for discrepancies between the techniques.
`
`KEYWORDS: pressurized metered-dose inhaler, im-
`pactor, time-of-flight, aerodynamic size distribution,
`aerosol measurement
`
`INTRODUCTION
`The particle size analysis of aerosols from pressurized
`metered-dose inhalers (pMDIs) by compendial proce-
`dures1,2 is typically undertaken using a multistage cas-
`cade impactor equipped with United States Pharma-
`copeia/European Pharmacopeia (USP/EP) induction
`port. This technique provides a direct link with the
`mass of therapeutically active pharmaceutical ingredi-
`ent (API) and particle aerodynamic size, which is ac-
`cepted as an indication of the likely deposition location
`within the respiratory tract.3 The recently introduced
`the Next Generation Pharmaceutical Impactor (NGI)
`(MSP, St Paul, MN4) was designed with the intent of
`improving the aerodynamic characteristics compared
`with the Andersen 8-Stage Cascade Impactor (ACI)
`(Thermo Andersen, Smyrna, GA) that is in widespread
`use for pMDI performance testing. The resulting im-
`paction-stage collection efficiency curves of the NGI at
`30 L/min5 are generally steeper than those obtained
`with the ACI,6 offering the prospect that the size frac-
`tionation process within the former will be more accu-
`rate. However, apart from a study involving prototype
`instruments,7 there is as yet almost no information to
`guide users as to the performance of the NGI with this
`class of inhaler. Cascade impaction is labor intensive
`whichever multistage impactor is used, even with aids
`
` 1
`
`R.J. Reynolds Vapor
`IPR2016-01268
`R.J. Reynolds Vapor v. Fontem
`Exhibit 1028-00001
`
`

`

`AAPS PharmSciTech 2003; 4 (4) Article 54 (http://www.aapspharmscitech.org).
`to speed up sample recovery.8 There is therefore a con-
`pactor. Each canister was shaken for 10 seconds and
`then primed by actuating 3 times to waste; then each of
`tinued interest in the development of more efficient
`5 actuations was delivered at 30-second intervals, with
`techniques that can be used particularly for early-stage
`product development.9 In the absence of a more rapid
`the mouthpiece of the inhaler coupled on axis with the
`entry to the induction port. Flow through the impactor
`multistage impactor-based technique, the use of so-
`was maintained until 30 seconds following the last ac-
`called ‘real-time’ aerodynamic particle size analyzers
`tuation. The impactor was subsequently disassembled
`based on the time-of-flight (TOF) principle has become
`quite commonplace.10 These instruments are capable of
`and the API recovered quantitatively from the induc-
`tion port, collection plates, and after filter, and then
`making a particle size measurement in typically less
`assayed by high performance liquid chromatography
`than a minute, depending on the concentration of the
`(HPLC)-UV spectrophotometry in accordance with
`aerosol that is sampled. However, TOF analyzers are
`established internal procedures. The size distribution
`susceptible to coincidence measurement problems
`from each of the canisters was determined using the
`when more than one particle is present in the measure-
`ment zone.11,12 Furthermore, the inability of at least one
`generic stage cut sizes supplied by the manufacturer, in
`accordance with compendial practice.1
`type of analyzer in this class, Aerosizer, (TSI, St Paul,
`MN) to discriminate between particles comprising API
`
`and those of excipient/surfactant has been shown to
`NGI
`result in significant bias when sizing the aerosol from a
`particular pMDI-produced suspension formulation.13
`The NGI measurements were made at 30.0 L/min ±
`More recently, however, studies with both
`the
`5%, also following the practice described for the ACI
`Aerosizer-LD14 and predecessor model 3320 Aerody-
`in the compendial method.1 The 304 stainless steel col-
`namic Particle Sizer (APS) aerosol spectrometer15
`lection cups were not coated with an adhesive agent,
`(TSI) have indicated that closer agreement with multi-
`based on previous experience using this impactor with
`stage impactor measurements may be possible for solu-
`pMDI-based aerosols.7 The NGI was used as supplied
`tion formulations where surfactant is absent. The APS
`for measurements with both chlorofluorocarbon
`is also supplied with the option of using a model 3306
`(CFC)-fluticasone propionate (Flovent, GSK Inc., Re-
`Single-Stage Impactor Inlet (SSI) (TSI), having a cut-
`search Triangle Park, NC) and CFC-sodium cromogly-
`point size of 4.7 µm aerodynamic diameter, to verify
`cate (Intal, Rhône-Poulenc Rorer Canada Inc., Mon-
`the magnitude of the so-called ‘respirable’ mass frac-
`tréal, QC, Canada), since the micro-orifice collector
`tion determined by the TOF analyzer.
`(MOC) acted as a substitute for a backup filter. How-
`ever, measurements made with a prototype instrument
`
`with Qvar had indicated that the MOC by itself might
`MATERIALS AND METHODS
`not have captured all of the extra-fine particles that
`penetrated beyond stage 7.7 An external filter unit
`Formulations
`(MSP) containing 2 layers of 934-AH glass microfiber
`Three pMDI-produced anti-asthmatic aerosols having
`was therefore connected to the outlet of the NGI for
`distinctly different particle size distribution properties
`measurements with this formulation. The operation of
`were evaluated (Table 1). Five canisters were chosen
`the pMDI canisters was as described for measurements
`at random from each of these formulations.
`by ACI.
`
`
`ACI
`APS and SSI
`Benchmark measurements were made using an alumi-
`The APS and SSI were operated together. The APS
`num ACI, sampling at 28.3 L/min ± 5%, following the
`counts particles as they pass individually through the
`procedure described in the USP.1 The ACI contained
`measurement zone where their aerodynamic size is de-
`uncoated glass collection plates with a backup glass
`termined, so it was necessary to transform the raw TOF
`microfiber filter (934-AH, Whatman, Clifton, NJ) lo-
`data to a mass-weighted size distribution using the pro-
`cated after the bottom impaction stage. In the case of
`prietary software provided (Aerosol Instrument Man-
`the measurements with hydrofluoroalkane (HFA)-
`ager, rev B [2002], TSI). The aerosol emitted from the
`beclomethasone dipropionate (Qvar) (3M Pharmaceu-
`inhaler was withdrawn at the nominal 28.3 L/min flow
`ticals, London, ON, Canada), 2 filters were used to-
`rate via a USP/EP induction port into the SSI, where
`gether in order to optimize collection of the small mass
`the incoming aerosol was sampled isokinetically at
`of extra-fine particles that penetrated beyond the im-
`
` 2
`
`R.J. Reynolds Vapor Exhibit 1028-00002
`
`

`

`AAPS PharmSciTech 2003; 4 (4) Article 54 (http://www.aapspharmscitech.org).
`Table 1. pMDI Produced Aerosols Evaluated by the Aerodynamic Particle Sizing Methods*
`Name
`Manufacturer
`Formulation Description
`Flovent-125
`GSK Inc (Canada)
`CFC-11/12 propellant mixture
`
`
`Lecithin surfactant
`
`
`125 µg/actuation fluticasone propionate†
`
`Intal-1 mg
`
`
`
`Rhône-Poulenc Rorer Inc (Canada) CFC-11/12 propellant mixture
`
`Sorbitan trioleate surfactant
`
`1000 µg/actuation sodium cromoglycate†
`
`Qvar-100
`
`
`
`
`
`3M Pharmaceuticals (Canada)
`
`
`
`HFA-134a propellant
`Ethanol cosolvent
`No surfactant
`100 µg/actuation beclomethasone dipropi-
`onate†
`*CFC indicates chlorofluorocarbon; HFA, hydrofluoroalkane; and pMDI, pressurized metered-dose inhaler.
`†Mass API/actuation expressed ex metering valve.
`
`similar [T. J. Beck, TSI Inc, November 2002 conversa-
`tion]. The operation of the pMDI canisters was again as
`described for the measurements using the multistage
`impactors.
`
`Interpretation of Data and Statistical Analysis
`The mass median aerodynamic diameter (MMAD) and
`geometric standard deviation (GSD), representing the
`measures of central tendency and spread, respectively,
`were used as metrics with which to compare the size
`distribution data. Since the central region (between
`16th and 84th percentiles) of the size distributions of all
`3 formulations obtained from the multistage impactors
`was in general well described by a log-normal distribu-
`tion function, the raw data were subjected to nonlinear
`regression analysis in accordance with the technique
`described by Thiel16 in order to establish values of
`MMAD without the need to interpolate. The APS pro-
`vides 43 size classes between 0.52 and 10.4 µm aero-
`dynamic diameter, so that error associated with interpo-
`lation between adjacent size classes to determine the
`MMAD was judged in this instance to be sufficiently
`small to be acceptable.
`FPF<4.7 µm, was also determined from the size distribu-
`tion data since this parameter is appropriate as a meas-
`ure of the therapeutically beneficial portion of the in-
`haled mass of anti-asthmatic medications capable of
`reaching the airways of the lower respiratory tract.17
`
`
`
`0.062 L/min (0.2% of the sample) directly to the APS
`(Figure 1).
`The remainder of the flow passed through the SSI. The
`portion of the mass entering this impactor contained in
`particles smaller than 4.7 µm aerodynamic diameter
`(defined as the fine particle or “respirable” fraction
`[FPF<4.7 µm]) was determined by HPLC-UV spectropho-
`tometric assay for the API collected on the after-filter
`of the impactor (containing 2 layers of 934-AH glass
`microfiber) and used to verify the equivalent result pre-
`sented by the TOF-based particle size measurements
`made using the APS. On this basis, FPF<4.7 µm could be
`determined as a percentage of the total mass entering
`the SSI in accordance with:
`
` 3
`
`(1)
`
`100
`
`
`
`
`
`)
`
`filter
`M
`+
`
`filter
`
`M
`
`M(
`
`stage
`
`
`
`FPF
`<
`
`mµ7.4
`
`=
`
`where Mstage and Mfilter are the masses of API that col-
`lect on the stage impaction plate and backup filter of
`this impactor, respectively.
`Where appropriate, a correction was applied to the
`APS-measured size distribution data to account for
`size-related losses in the sampling system. This correc-
`tion was based on the 100:1 size-efficiency relationship
`obtained for the Aerosol Diluter (model 3302A, TSI),
`also available for use with the APS, on the basis that
`the capillary dimensions and aerosol pathway from the
`isokinetic nozzle to the exit of the impactor inlet were
`
`R.J. Reynolds Vapor Exhibit 1028-00003
`
`

`

`AAPS PharmSciTech 2003; 4 (4) Article 54 (http://www.aapspharmscitech.org).
`
`Figure 1. Schematic of model 3306 showing flow pathways to the single-stage impactor and APS. (Cour-
`tesy TSI Inc).
`
`
`
`
`This parameter was obtained directly from the size dis-
`tributions measured by both ACI and APS as both in-
`struments have size class limits that correspond exactly
`to 4.7 µm aerodynamic diameter. FPF<4.7 µm could also
`be determined directly from the SSI, since its cut size is
`fixed at 4.7 µm aerodynamic diameter. FPF<4.7 µm was
`estimated by linear interpolation for the NGI since
`
`stages 2 and 3 have cut sizes of 6.4 and 4.0 µm aerody-
`namic diameter, respectively, at 30 L/min.
`Statistical interpretation of the data derived from the
`size distributions obtained by the various procedures
`was undertaken using appropriate tests of significance
`(SigmaStat, version 2.3, SPSS Science, Chicago, IL).
`Differences were deemed significant when P < .05.
`
` 4
`
`R.J. Reynolds Vapor Exhibit 1028-00004
`
`

`

`AAPS PharmSciTech 2003; 4 (4) Article 54 (http://www.aapspharmscitech.org).
`Values of the reported performance metrics represent
`mean ± SD based on 5 replicate measurements unless
`otherwise stated.
`
`RESULTS AND DISCUSSION
`The choice of the ACI as the benchmark device for the
`present study reflects the widespread use of this impac-
`tor for the measurement of pharmaceutical aerosols by
`the compendial procedure.1 The results from this study
`do not enable any claim to be made in terms of the ac-
`curacy of this impactor in comparison with the other
`techniques.
`Mass recovery of API was within ± 20% of label claim
`for the measurements with both ACI and NGI. The
`mean mass loading of the NGI, based on 5 actuations
`per measurement and considering only the mass that
`penetrated beyond the induction port to the impactor,
`was substantially greater for Intal (1443 µg) compared
`with either Qvar (227 µg) or Flovent (246 µg). Similar
`total mass loading data (not shown) were obtained for
`the ACI.
`Comparative size distributions for the ACI, NGI, and
`APS for Flovent, Intal, and Qvar are summarized on a
`cumulative mass-weighted basis in Figures 2, 3, and 4,
`respectively, using log-probability scaling.
`Only minor differences were observed in GSD values
`between the 3 measurement techniques for Qvar and
`Intal (Table 2), and GSDs for Flovent aerosols were
`equivalent (P = .87). No technique, therefore, consis-
`tently produced size-distribution data that were consis-
`tently less or more disperse than data obtained by the
`other 2 instruments. However, although values of
`MMAD for Intal determined by either of the multistage
`impactors (4.3 ± 0.19 µm [NGI], 4.2 ± 0.13 µm [ACI])
`were comparable (unpaired t test, P = .29), the NGI-
`measured MMAD for Flovent (2.0 ± 0.05 µm) was
`significantly finer than that obtained by ACI (2.8 ±
`0.07 µm) (P < .001). The ACI-based MMAD was,
`however, within the range from 2.4 to 2.8 µm reported
`by Cripps et al for this formulation, also using this type
`of impactor.18
`Overlap of the collection efficiency curves of neighbor-
`ing stages of either impactor is an unlikely cause of the
`observed differences between MMAD values obtained
`from the multistage impactors for Flovent, as the effect
`is reported to be small below stage 2 based on a previ-
`ously published calibration of an ACI,6 and should be
`even less apparent with the NGI in view of its sharp
`and well-separated stage collection efficiency curves.5
`
`
`Figure 4. Comparison of ACI-, NGI-, and APS-
`measured size distributions for Qvar.
`
`
` 5
`
`Figure 2. Comparison of ACI-, NGI-, and APS-
`measured size distributions for Flovent.
`
`
`
`Figure 3. Comparison of ACI-, NGI-, and APS-
`measured size distributions for Intal.
`
`
`R.J. Reynolds Vapor Exhibit 1028-00005
`
`

`

`AAPS PharmSciTech 2003; 4 (4) Article 54 (http://www.aapspharmscitech.org).
`Table 2. Size Distribution Parameters Obtained From the ACI, NGI, and APS*
`
`Flovent
`Intal
`
`Qvar
`
`
`
`MMAD (µm)
`
`GSD
`
`MMAD (µm)
`
`GSD
`
`MMAD (µm)
`
`GSD
`
`1.84 ± 0.06
`1.2 ± 0.02
`1.63 ± 0.04
`4.2 ± 0.13
`1.56 ± 0.03
`2.8 ± 0.07
`ACI†
`1.82 ± 0.03
`0.9 ± 0.03
`1.75 ± 0.04
`4.3 ± 0.19
`1.55 ± 0.07
`2.0 ± 0.05
`NGI†
`1.74 ± 0.05
`1.0 ± 0.03
`1.66 ± 0.04‡
`3.2 ± 0.02‡
`1.57 ± 0.02
`1.8 ± 0.07
`APS
`*ACI indicates Andersen 8-stage cascade impactor; APS, aerodynamic particle sizer; GSD, geometric standard deviation; MMAD, mass
`median aerodynamic diameter; and NGI, next generation pharmaceutical impactor. Except where indicated, n = 5 replicates/technique.
`†Size distribution parameters are based on mass penetrating the induction port and entering the impactor.
`‡4 replicate measurements.
`
`
`
`There are at least 2 other possibilities to consider, and
`both potential causes may have contributed to the ob-
`served behavior. Kamiya et al19 have recently reported
`that the NGI, if used with uncoated collection cups,
`undersized a CFC-suspension formulation (Vanceril,
`Schering-Plough, Kenilworth, NJ) that has similar size
`distribution characteristics to Flovent. They obtained
`close agreement between NGI- and ACI-measured data
`when their cups were precoated with silicone oil to im-
`prove particle adhesion. The NGI does appear to be at
`greater risk than the ACI of bias caused by this effect.
`Stage Reynolds numbers, which govern particle veloc-
`ity and therefore kinetic energy at the point of collec-
`tion in an impactor,20 are in the range from 324 to 2938
`at 30 L/min for the NGI, as a direct consequence of
`optimizing its aerodynamic performance,4 whereas the
`equivalent range for the ACI is from 110 to 782.20 At
`first sight, the agreement achieved between NGI- and
`ACI-measured size distributions for Intal is difficult to
`reconcile. However, although the sodium cromoglycate
`particles are in general larger than those formed from
`Flovent, they are hygroscopic,21 likely making their
`surfaces tacky under the conditions of the present
`study, and therefore more prone to adhere to an un-
`coated surface. In view of these observations, it may be
`prudent to coat the collection cups of the NGI with an
`adhesive agent to be sure that particle bounce does not
`occur, unless it can be demonstrated that this phe-
`nomenon is not occurring to a significant extent with a
`given formulation.
`It is also possible that propellant evaporation was not
`fully complete by the time that sampling of particles
`produced from Flovent took place. Under these cir-
`cumstances, the size distribution measured by the im-
`pactor having the larger internal volume (1000 mL,
`NGI; 450 mL, ACI) might be expected to be finer as a
`result of more complete evaporation having taken place
`by the time that the particles were collected. The boil-
`
`ing point of CFC-11 is 23.8°C at 101.3 kPa, so that at
`room ambient conditions close to 22°C, evaporation of
`this component would be expected to be relatively slow
`compared with that of CFC-12 (boiling point -26.1°C).
`However, published data on this effect are scant.
`Morén22 commented that the initial flashing of liquid
`propellant to vapor as pressure is relieved upon actua-
`tion is so rapid that heat required for the change of
`phase is taken from the liquid remaining, therefore re-
`sulting in cooling of the droplets. Further evaporation
`then occurs as energy is acquired from the surrounding
`air molecules; this evaporation is slow compared with
`the flashing process. Morén and Anderson,23 in a study
`involving a formulation containing terbutalene sulfate
`in a CFC-11:12:114 mixture, reported MMAD values
`that decreased from 43 µm immediately after actuation
`to 14 µm at a distance of 10 cm from the canister. On
`this basis, aerosol transport beyond the induction port
`might therefore have taken place before propellant
`evaporation was complete. However, precise evapora-
`tion behavior will depend on actuator orifice diameter,
`the mass of formulation metered per actuation, and
`aerosol pathway into the impactor, as well as physical
`properties of the propellant,24 so that data from studies,
`such as that of Morén and Anderson can only provide a
`general indication of the magnitude of this effect.
`The agreement between measurements by both impac-
`tors for Intal in the present study is also explicable in
`terms of propellant evaporation behavior if the kinetics
`of evaporation of propellant from droplets containing a
`large mass concentration of solids (as is the case with
`Intal) approached equilibrium more rapidly. Further-
`more, the relatively large particles produced by Intal
`compared with those formed from Flovent, travel a
`shorter distance into either impactor before being col-
`lected, so that differences in internal volumes between
`the NGI and ACI would be expected to be less impor-
`tant.
`
` 6
`
`R.J. Reynolds Vapor Exhibit 1028-00006
`
`

`

`AAPS PharmSciTech 2003; 4 (4) Article 54 (http://www.aapspharmscitech.org).
`The APS undersized aerosols from both Flovent and
`Intal, compared with either multistage impactor. Thus,
`the MMAD measured by the APS was 1.8 ± 0.07 µm
`for Flovent and 3.2 ± 0.02 µm for Intal (1-way analysis
`of variance [ANOVA] for each formulation, P < .05).
`The cause is believed to be the presence of surfactant
`particles that are formed together with particles com-
`prising API with both formulations. It is pertinent that
`in a previous study using an Aerosizer TOF aerody-
`namic particle size analyzer, the Aerosizer-measured
`MMAD of a suspension formulation containing
`budesonide with sorbitan trioleate surfactant was found
`to be 2.4 ± 0.2 µm, compared with 3.9 ± 0.1 µm by
`ACI.13 The surfactant particles were observed by mi-
`croscopy to have collected further into the impactor
`(finer sizes) compared with particles of API, and the
`Aerosizer-measured size distributions were shown to
`have included both surfactant and API particles. The
`APS, like the Aerosizer, cannot discriminate between
`particles of surfactant and API, since no assay for drug
`substance is undertaken.
`In the case of Flovent, the ACI-measured FPF<4.7 µm
`(88.8% ± 2.9%) was slightly lower than 93.4% ± 0.7%
`obtained with the NGI (P < .001) (Table 3), consistent
`with the larger MMAD obtained by the ACI. FPF<4.7 µm
`from the SSI (used in conjunction with the APS [82.9%
`± 2.1%]) was only slightly smaller than the ACI-
`measured value (P =
`.006). However, the SSI-
`determined FPF<4.7 µm was considerably less than the
`97.9% ± 1.2% obtained using the APS (P < .001), as a
`consequence of the tendency for the APS to undersize
`compared with the impactors. For Intal, FPF<4.7 µm
`measured by the SSI (46.4% ± 2.4%) was also mark-
`edly smaller than 78.8% ± 2.1% obtained by the APS
`(P <.001), and also slightly lower than corresponding
`values determined by the multistage impactors (60.2%
`± 2.8% [ACI]; 52.0% ± 2.9% [NGI]) (P < .05). This
`slight discrepancy between SSI- and multistage impac-
`tor data can be explained if the CFC-11 propellant
`evaporation was incomplete by the time that the parti-
`cles entered the SSI, whose internal volume is less than
`200 mL [G. Pence, TSI Inc, letter, February 2003].
`In contrast to the behavior with Flovent and Intal, the
`APS-measured MMAD (1.0 ± 0.03 µm) for Qvar,
`which is a solution formulation containing no surfac-
`tant, was located between the corresponding values
`from the NGI and ACI, and the difference between this
`value and the NGI-measured MMAD was barely sig-
`nificant (P = .011). Hence, FPF<4.7 µm from both the
`APS and the multistage impactors were in close
`agreement (98.0% ± 0.5% [ACI]; 96.7% ± 0.7%
`[NGI]; 96.4% ± 2.5% [APS]) (P = .18). The steeper
`
`
`Figure 5. Effect of correction for inlet sampling effi-
`ciency on APS-based size distribution measurements
`for Intal.
`
`NGI-measured particle size distributions for Qvar were
`also finer than those obtained by the ACI (MMAD =
`0.9 ± 0.03 [NGI]; 1.2 ± 0.02 [ACI]) (P < .001). Again
`the ACI-based MMAD was in close agreement with
`1.1 µm reported previously for this formulation, also
`using this type of impactor.25 In the case of this formu-
`lation, the finer MMAD measured by the NGI com-
`pared with ACI can also be explained in terms of dif-
`fering evaporation behavior. However, in this instance,
`the ethanol solubilizer (boiling point 78°C) is more
`likely than HFA-134a propellant (boiling point -
`26.5°C) to have been incompletely evaporated when
`the particles were collected. This explanation is sup-
`ported by data from a study by Gupta et al,26 who
`demonstrated that ethanol evaporation is incomplete at
`the distal end of a USP induction port, working with a
`range of solution formulations similar in composition
`to Qvar.
`It should be noted that the APS-measured size distribu-
`tions were not corrected for particle counting efficiency
`as a function of particle size, since Peters and Leith27
`recently indicated that counting efficiency is size inde-
`pendent in the range of 45% to 60% between 0.7- and
`4-µm aerodynamic diameter for the model 3321 APS.
`However, a correction was made for the sampling effi-
`ciency of the SSI (see Materials and Methods section).
`The effect of this correction on the reported size distri-
`bution data was found to be negligible, except for Intal
`(Figure 5), since the majority of the particles sampled
`by the APS for Flovent and Qvar were finer than 4-µm
`aerodynamic diameter, where the efficiency of the
`Aerosol Diluter is close to 100%.28
`
` 7
`
`R.J. Reynolds Vapor Exhibit 1028-00007
`
`

`

`AAPS PharmSciTech 2003; 4 (4) Article 54 (http://www.aapspharmscitech.org).
`Table 3. Values of FPF<4.7µm (%) Obtained From the ACI, NGI, and APS*
`Qvar
`
`Flovent
`Intal
`98.0 ± 0.5
`88.8 ± 2.9
`60.2 ± 2.8
`ACI†
`96.7 ± 0.7
`93.4 ± 0.7
`52.0 ± 2.9
`NGI†
`96.4 ± 2.5
`97.9 ± 1.2
`78.8 ± 2.1‡
`APS
`67.1 ± 4.1
`82.9 ± 2.1
`46.4 ± 2.4
`3306 Impactor
`*ACI indicates Andersen 8-stage cascade impactor; APS, aerodynamic particle sizer; FPF<4.7 µm, fine particle fraction <4.7 µm aerody-
`namic diameter; and NGI, next generation pharmaceutical impactor. Except where indicated, n = 5 replicates/technique.
`†Values are based on mass penetrating the induction port and entering the impactor.
`‡ Based on 4-replicates.
`
`
`
`slope of the APS-measured data for particles finer than
`0.8 µm aerodynamic diameter (Figure 4) can be ex-
`plained by loss of sensitivity of the TOF detection sys-
`tem, which reaches its limit of detection at about 0.5
`µm aerodynamic diameter.29 However, FPF<4.7 µm de-
`termined by the SSI (67.1% ± 4.1%) was surprisingly
`much smaller than any of the other values for this for-
`mulation, which ranged from 96% to 98% (P < .001)
`(Table 3). Similar behavior was reported by Gupta et
`al26 and explained in terms of incomplete ethanol
`evaporation in the SSI. A solution might be to extend
`the passageway to the SSI to match more closely the
`aerosol transit time with that for the APS, as was done
`by Gupta et al, who found that as much as 40 cm of
`inlet extension was needed to achieve good agreement
`between techniques. Such a change would also be ex-
`pected to improve the agreement between SSI- and
`multistage impactor-measured FPF<4.7 µm for formula-
`tions, such as Flovent and Intal that contain relatively
`low volatile CFC-11 propellant. However, it will be
`necessary to be careful not to introduce additional sur-
`faces for impaction if a design change of this nature is
`contemplated. An alternative strategy, avoiding such
`problems, might be to heat the existing transfer channel
`so that the ethanol/propellant is more rapidly evapo-
`rated. Whichever solution is implemented, some flexi-
`bility will be needed for the user to set up the aerosol
`transport conditions in the SSI and APS on a formula-
`tion-by-formulation basis, given the variety of volatile
`species that may be present within the range of pMDIs
`that are available.
`
`CONCLUSION
`These in vitro comparisons indicate how 3 pMDI-
`generated formulations behaved in laboratory-based
`sampling systems. Discrepancies in size distributions
`were identified in data from 2 different designs of cas-
`cade impactor operated under near equivalent condi-
`
`tions. Aerosols from Flovent and Intal were undersized
`by the APS, most likely because of its inability to dis-
`criminate between particles containing API and finer
`surfactant particles. Further work is merited to resolve
`whether apparent undersizing observed with the NGI
`with both Flovent and Qvar is caused either by particle
`bounce and/or incomplete evaporation of volatile con-
`stituents. The SSI should be modified to ensure compa-
`rable evaporation of volatiles to that obtained in the
`APS, if formulations such as Qvar are to be sized accu-
`rately. It is strongly recommended that both NGI and
`APS/SSI data be evaluated on a formulation-by-
`formulation basis in relation to the large database that
`already exists for ACI-based measurements.
`
`ACKNOWLEDGEMENTS
`The authors acknowledge the support of Dr. Daryl
`Roberts, MSP Corp, St Paul, MN, for advice and assis-
`tance with the modifications to the NGI, and Mr. Greg
`Pence and Mr. Tyler Beck, TSI Inc, St Paul, MN, for
`the loan of a model 3321 APS with model 3306 Impac-
`tor Inlet and advice on use and interpretation of data.
`
`
`REFERENCES
`1. USP 26-NF 21. Chapter 601-Physical tests and determinations:
`aerosols. United States Pharmacopeia. Rockville, MD: United
`States Pharmacopeial Convention; 2003:2105-2123.
`2. European Pharmacopeia. Section 2.9.18-Preparations for inha-
`lation: aerodynamic assessment of fine particles. European Phar-
`macopeia. 3rd ed. [Suppl 2001]. Strasbourg, France: Council of
`Europe; 2002:113-124.
`3. Rudolph G, Kobrich R, Stahlhofen W. Modeling and algebraic
`formulation of regional aerosol deposition in man. J Aerosol Sci.
`1990;21(suppl 1):306-406.
`4. Marple VA, Roberts DL, Romay FJ, Miller NC, Truman KG,
`Van Oort M, Olsson B, Holroyd MJ, Mitchell JP, Hochrainer D.
`Next generation pharmaceutical impactor. Part 1: Design. J Aero-
`sol Med. 2003;16:283-299.
`
` 8
`
`R.J. Reynolds Vapor Exhibit 1028-00008
`
`

`

`AAPS PharmSciTech 2003; 4 (4) Article 54 (http://www.aapspharmscitech.org).
`17. Dolovich MB. Aerosol delivery devices and airways/lung
`5. Marple VA, Olson BA, Santhanakrishnan K, Mitchell JP,
`Murray SC, Hudson-Curtis BL. Next generation pharmaceutical
`deposition. In: Schleimer RP, O’Byrne P, Szeffler S, Bratsand R,
`impactor. Part 2: Archival calibration. J Aerosol Med.
`eds. Inhaled Steroids in Asthma. New York, NY: Marcel Dekker;
`2003;16:301-324.
`2001:169-210.
`18. Cripps A, Riebe M, Schulze M, Woodhouse R. Pharmaceuti-
`6. Mitchell JP, Costa PA, Waters S. An assessment of an Ander-
`sen mark-II cascade impactor. J Aerosol Sci. 1988;19(2):213-231.
`cal transition to non-CFC pressurized metered dose inhalers. Res-
`pir Med. 2000;94(suppl B):3-9.
`7. Mitchell JP. The next generation impactor (NGI): results from
`the evaluation of prototype instruments with pressurized metered
`19. Kamiya A, Sakagami M, Hindle M, Byron PR. Particle sizing
`dose inhaler (pMDI)-based formulations. In: Drug Delivery to the
`with the next generation impactor: a study of Vanceril™ metered
`Lungs – XI. London, UK: Aerosol Society; 2000:223-226.
`dose inhaler. J Aerosol Med. 2003;16(2):216.
`8. Miller NC, Roberts DL, Marple VA. The ‘Service Head’ ap-
`20. Marple VA, Olson BA, Miller NC. The role of inertial particle
`proach to automating the next generation pharmaceutical impac-
`collectors in evaluating pharmaceutical aerosol delivery systems. J
`tor: proof of concept. In: Dalby RN, Byron PR, Peart J, Farr SJ,
`Aerosol Med. 1998;11(suppl 1):139-153.
`eds. Respiratory Drug Delivery VIII. Raleigh, NC: Davis Hor-
`21. Chew NYK, Bagster DF, Chan H-K. Effect of particle size, air
`wood International; 2002:521-523.
`flow and inhaler device on the aerosolisation of disodium cromo-
`

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket