throbber
Downloaded from journalpdaorg on May 31, 2012
`
`
`
`EDA
`PDA Journal
`of Pharmaceutical Science and Technology W
`
`
`
`
`Variability in Syringe Components and its Impact on
`Functionality of Delivery Systems
`Nitin Rathore, Pratik Pranay, Bruce Eu, et al.
`
`PDA J Pharm Sci and Tech 2011, 65 468-480
`Access the most recent version at doi:10.5731/pdajpst2011.00785
`
`REGITCOOO72194
`
`Regeneron Exhibit 1087.001
`
`

`

`Downloaded from journalpdaorg on May 31, 2012
`
`Variability in Syringe Components and its Impact on
`Functionality of Delivery Systems
`
`NITIN RATHORE“, PRATIK PRANAYZ, BRUCE :U‘, WENCHANG .JI‘, and ED WALLS1
`
`[Drug Product and Device Development, Amgen, Thousand Oaks, CA and 2Department of Chemical Engineering,
`University of WisconsiniMadz'son, Madison. WI ©PDA, Inc. 2011
`
`ABSTRA CT: Prefilled syringes and autoinjectors are becoming increasingly common for parenteral drug administra—
`tion primarily due to the convenience they offer to the patients. Successful commercialization of such delivery
`systems requires thorough characterization of individual components. Complete understanding of various sources of
`variability and their ranking is essential for robust device design. In this work, we studied the impact of variability
`in various primary container and device components on the delivery forces associated with syringe injection. More
`specifically, the effects of barrel size, needle size, autoinjector spring force. and frictional forces have been evaluated.
`An analytical model based on underlying physics is developed that can be used to fully characterize the design space
`for a product delivery system.
`
`KEYWORDS: Break-loose forces, Extrusion force, Device, Autoinjcctor, Prefilled syringe, Delivery forces
`
`LAY ABSTRACT: Use of prefilled syringes (syringes prefilled with active drug) is becoming increasingly common for
`injectable drugs. Compared to vials, prefilled syringes offer higher dose accuracy and ease of use due to fewer steps
`required for dosage. Convenience to end users can be further enhanced through the use of prefilled syringes in
`combination with delivery devices such as autoinjectors. These devices allow patients to self—administer the drug by
`following simple steps such as pressing a button. These autoinjectors are often spring—loaded and are designed to keep
`the needle tip shielded prior to injection. Because the needle is not visible to the user, such autoinjectors are perceived
`to be less invasive than syringes and help the patient overcome the hesitation associated with self—administration.
`In order to successfully develop and market such delivery devices, we need to perform an in—depth analysis of the
`components that come into play during the activation of the device and dose delivery. Typically, an autoinjector is
`activated by the press of a button that releases a compressed spring; the spring relaxes and provides the driving force
`to push the drug out of the syringe and into the site of administration. Complete understanding of the spring force,
`syringe barrel dimensions, needle size. and drug product properties is essential for robust device design.
`it is equally important to estimate the extent of variability that exists in these components and the resulting impact
`it could have on the performance of the device. In this work, we studied the impact of variability in syringe and device
`components on the delivery forces associated with syringe injection. More specifically, the effect of barrel size, needle
`size, autoinjector spring force, and frictional forces has been evaluated. An analytical model based on underlying
`physics is developed that can be used to predict the functionality of the autoinjector.
`
`Introduction
`
`The last decade has witnessed an increase in the pop—
`ularity and sales of prefilled syringes with an annual
`growth rate of 20% in the U.S. market (1). The pri—
`mary factors driving the growth include ease of ad—
`
`
`
`* Corresponding Author: Drug Product and Device
`Development, Amgen, One Amgen Center Dr., MS
`30W—3-A, Thousand Oaks, CA 91320; Phone 805-
`313—6393; E—mail: nrathore@amgen.com
`
`ministration and added convenience for health care
`
`workers and patients (1, 2). Compared to vial config-
`uration, a higher accuracy can be achieved with pre-
`filled syringes and fewer steps are required for dosage.
`An added benefit is the reduced overfill amount due to
`
`significantly lower hold—11p volumes associated with
`syringes. Errors in dosage, and risk of misidentifica—
`tion and contamination, are also minimized. Plastic
`
`prefillable syringes made of cyclic olefins are now
`available as an alternative to glass syringes (3). Con-
`venience to end users and market advantage can fur-
`ther be boosted through the use of delivery devices.
`Delivery systems that are preferred by the patients and
`
`468
`
`PDA Journal of Pharmaceutical Science and Technology
`
`REGITCOOO72195
`
`Regeneron Exhibit 1087.002
`
`

`

`Downloaded from journalpdaorg on May 31, 2012
`
`perceived to be less invasive than syringes (4) will
`provide commercial advantage to the drug manufae~
`turer. Novel delivery systems for commercial products
`also offer a mechanism to maintain the competitive
`edge in the marketplace (5).
`
`Successful commercialization of prcfillcd syringe eon-
`figurations and autoinjectors requires complete under-
`standing of the mechanism of delivery and the param—
`eters contributing to the delivery forces and injection
`time. The delivery force is attributed to the break-
`loose force (initial force required to set the plunger in
`motion) and the extrusion force needed to sustain the
`plunger movement by overcoming the hydrodynamic
`pressure and the frictional forces. Several factors con—
`tribute to these forces, including but not limited to:
`
`O Barrel siliconization, which primarily affects the
`frictional forces
`
`'
`
`O
`
`'
`
`'
`
`Syringe geometry, including barrel size and needle
`gauge, which primarily affects the force due to
`hydrodynamic pressure drop
`
`Syringe type, such as Siliconized glass or plastic
`
`Stopper type and geometry
`
`including its interaction with
`Product attributes,
`the barrel surface and its rheological properties
`
`' Driving forces, such as the spring for mechanical
`autoinjectors
`
`'
`
`'
`
`Injection volume and time
`
`Subcutaneous resistance
`
`In order to design a robust product presentation, it is
`important
`to understand the role of each of these
`components, estimate their inherent variability, and
`calculate the resulting impact on injection force or
`time. The objective of this study is to characterize and
`measure the effect of variability in components asso-
`ciated with a syringe delivery system, such as syringe
`barrel size, needle size, friction forces, and spring
`characteristics of the autoinjector. The role of product
`properties and its interaction with the syringe surface
`is equally critical and has been evaluated under a
`separate study. Results from that study will be pub-
`lished in a separate article. Subcutaneous resistance is
`also expected to increase the delivery forces; however,
`
`Vol. 65, No. 5, September—October 2011
`
`TABLE I
`
`List of Syringe Types and Lots Used in This
`Study
`
`Number
`
`
`
`
`Syringe
`of Lots
`
`Vendor
`Studied
`Syringe Type
`I"
`I"
`'“i
`
`Vendor l
`3 lots
`Siliconized glass
`Vendor 2
`3 lots
`Siliconized glass
`j
`Vendor 3
`2 lots
`r Siliconized glass
`
`L Vendor4
`l
`lot
`Plastic
`
`4
`
`the impact of interstitial pressure is outside the scope
`of this work. The measurements of extrusion forces
`
`are performed using Instron. a material testing system.
`A predictive model based on the Hagen—Poiseuille
`equation has been developed to understand the flow
`behavior of drug through the delivery systems and to
`help identify malfunctions and failure points associ-
`ated with the delivery system. The mechanistic model
`helps to identify the key process parameters, assess
`their importance, and predict the impact they would have
`on the extrusion force or injection time variability.
`
`Materials and Methods
`
`Siliconized glass syringes and plastic syringes pro—
`cured from different vendors (see Table I) were used
`in this study. Plunger stoppers from two different
`vendors were also evaluated for siliconized glass sy-
`ringes.
`
`Force measurements for syringes were performed us-
`ing Instron, a material testing system. A load cell of
`500 N was used to drive the syringe plunger at a
`constant crosshead speed while measuring the result-
`ing force on the plunger (repeatability of :0.25% of
`reading over a range of 0.4% to 100% of capacity). A
`schematic of the instrument
`is shown in Figure 1.
`Variation in needle size is measured by a syringe flow
`rate fixture which measures the pressure drop for a
`liquid (water) flowing across the syringe barrel and
`needle at a constant flow rate. The set up consists of a
`pump connected to a water reservoir and a pressure
`sensor. The pump discharges water at constant flow
`rate in the capillary, and the sensor measures the
`corresponding pressure drop that is representative of
`the effective internal radius of the needle. Variation in
`
`barrel size of the syringes is measured by the barrel
`bore internal diameter (ID) gauge. it is first calibrated
`using a barrel of known ID. The instrument is then
`
`469
`
`REGITCOOO72196
`
`Regeneron Exhibit 1087.003
`
`

`

`Downloaded from journalpdacrg on May 31, 2012
`
`
`
`Figure 1
`
`Picture of the lnstron system used for measuring
`extrusion forces.
`
`used to measure the barrel size of different syringes at
`different depths along the syringe axis.
`
`Theory
`
`The system under consideration is fluid flow through a
`prefilled syringe. The syringe consists of a needle of
`length Ln and mean effective internal radius rn at—
`tached to the barrel of mean effective internal radius
`
`rb. The syringes are filled to a specified volume and
`stoppered using an automated stopper placement unit.
`The stopper holds the end of the plunger rod through
`which a force Ftotal is applied in order to drive the fluid
`with a plunger speed \7 (linear speed in length over
`time dimensions).
`
`Break-loose force refers to the maximum force re—
`quired to set the plunger into motion. Extrusion force
`is the total force required to sustain the plunger rod in
`motion While maintaining the desired flow rate of the
`liquid through the needle. This study characterizes the
`
`total extrusion force associated with delivery of a
`product through syringe injection.
`
`Figure 2 shows the schematic of a syringe system. The
`inner surface of the glass barrel
`is lubricated with
`silicone oil as shown in the figure. The force balance
`on the stopper at any time during injection gives
`
`ham 2 Tammi + Thydrmiynamiv
`
`(1)
`
`where Fmta, is the total force needed for driving the
`plunger (also referred to as extrusion force), meon is
`the friction force between the stopper and the syringe
`
`wall, and Fhwmdymmjc is the hydrodynamic force re-
`quired to drive the fluid out of the needle. The details
`of these forces are discussed in the following sections.
`
`A. Friction Force
`
`The friction force arises from the interaction between
`
`the walls of the stopper and the barrel. The inner
`surface of the glass syringe is lubricated (siliconizcd)
`with a thin layer of silicone oil as shown in Figure 2.
`The friction force thus results from the glass~silicone
`oil—stopper interaction. Using the lubrication approxi-
`mation, and assuming a uniform silicone oil layer on
`the inner wall of the barrel, the relation between the
`
`friction force and the injection speed is
`/
`
`__
`Ffiictioti _
`
`zwl’kmlrlilslt)
`.‘
`Oil
`d
`
`\
`
`ppci
`
`'
`
`,
`; __
`V _ 1(1'V
`
`(2)
`
`where no“ is the viscosity of lubricating oil, do” is the
`
`thickness of lubrication layer, ismpper is the length of
`the stopper in contact with glass, and v is the injection
`speed (linear piston speed with dimensions of length
`over time). Equation 2 shows that there is a linear
`dependence of the friction force on the injection speed
`and KC is the constant of proportionality for a given
`
`
`
`
`
`
`
`
`
`Figure 2
`
`stopper
`
`A schematic of the various components and forces associated with the syringe delivery system. The figure also
`shows a schematic of the lubrication of the syringe wall with silicone oil.
`
`470
`
`PDA Journal of Pharmaceutical Science and Technology
`
`REGITCOOO72197
`
`Regeneron Exhibit 1087.004
`
`

`

`Downloaded from journalpdacrg on May 31, 2012
`
`dence of the hydrodynamic force on the injection
`speed. Equations 2 and 4 can be combined to give the
`total extrusion force associated with syringe delivery:
`
`I:total _
`
`/21Tp’oilrblstopocr)_
`Cl
`V _
`
`oil
`
`\
`
`(SWMLnr:\)_
`D
`4
`V‘
`1‘
`
`/
`
`(5)
`
`C. Hydrodynamic Force for Non—Newtonian Fluids
`
`The flow of non—Newtonian fluids is more complex
`due to the fact that their viscosity is not constant with
`the shear rate. A power law model is most commonly
`applied to represent the viscosity for such fluids:
`
`7w = KW)”
`
`and
`
`M = Kb)“
`
`(6)
`
`the wall or barrel
`where TW is the shear stress at
`surfaeet y is the shear rate and n is the power law
`index (where n '— 1 represents a Newtonian fluid), K is
`the defined as the flow consistency index, and p. is the
`apparent Viscosity.
`
`For non—Newtonian fluids,
`
`the relation between the
`
`pressure drop AP required to drive the fluid at flow
`rate Q in a cylindrical channel of radius r and length L
`can be derived by solving the Navier—Stokes equation
`for a flow in a cylinder (6). Neglecting the pressure
`across the barrel, the hydrodynamic force required to
`drive a non—Newtonian fluid with an injection speed \7
`can be derived as
`
`thdrodynamrc
`
`
`(3n + l\ ” 2*7Klinrfi"+2
`n
`l
`
`3u+l
`
`<1l
`
`_
`
`— KhV"-
`
`(7)
`
`It should be noted that while the hydrodynamic force
`was linear with injection speed for the case of New—
`tonian fluids, it has a non—linear dependence on injec—
`tion speed for non-Newtonian fluids. The total extru-
`sion force can then be estimated by adding the friction
`force to the hydrodynamic component:
`
`thickness of silicone oil. The friction force would
`
`increase with injection speed due to the increase in
`velocity gradient within the lubrication layer. Vari—
`ability in friction force could arise due to non—unifor—
`mity in the thickness of the silicone oil layer on the
`inside surface of the barrel, as well as variations in the
`
`geometry of the barrel and stopper. Protein—barrel
`interactions could further affect the friction force,
`
`B. Hydrodynamic Force for Newtonian FlllldS
`
`The hydrodynamic force results from the pressure
`drop required to drive the fluid out of the syringe. For
`Newtonian fluids, the relationship between the pres—
`sure drop AP and the volumetric flow rate Q (units:
`volume/time) can be obtained using the Hagen—Poi—
`seuille law as
`
`
`8 L
`AP: ”4Q
`171‘
`
`(3)
`
`where it is the viscosity of fluid, r is the radius, and L
`is the length of the cylindrical channel. The equation
`assumes laminar flow (Re < 2300) for an incompress-
`ible liquid though a channel of constant cross section
`diameter of 2r. For a 27 G syringe needle and 1 mL
`syringe barrel used in this study, a plunger speed of
`304.8 mm/min corresponds to a Reynolds number of
`less than lOO. Assuming no interference from the glue
`used in producing a staked needle syringe,
`the total
`hydrodynamic force associated with flow in a syringe
`will depend on the pressure drop across the barrel and
`needle. Equation 3 shows that for constant flow rate Q,
`AP ~ r74. In the syringe system, the radius of the
`barrel is much larger than the radius of the needle
`(rb/rn E 30). As a result, the pressure drop across the
`barrel
`is negligible when compared to the pressure
`drop across the needle (~0(10’6)). There is also an
`entry loss when the fluid enters into a constriction, but
`its magnitude is much smaller than the pressure drop
`across the needle (APIOSS ~ pv2/2 ~ 00077)). Ne—
`glecting the pressure drop across the syringe barrel
`and the entry loss, the hydrodynamic force at a given
`temperature can be estimated from eq 3 as
`
`Fliydtodynamic = (—4. V = KhV
`
`8’rruLnr: _
`1‘H
`
`_
`
`(4)
`
`.
`/
`_ Zfip’oillhlstopoe» __
`Ftotal "7 d—)V
`\
`oil
`
`
`+
`
`(311 'l'
`
`\n
`
`l)“21’rKL“rfi“+2
`
`7
`
`V“~
`
`(8)
`
`
`
`n+1n
`
`r
`
`where Kh is a constant that depends on syringe size
`and fluid properties. Variation in operating tempera-
`ture would affect the solution viscosity and the hydro-
`dynamic force. Equation 4 shows the linear depen—
`
`D. Injection Time Calculation for Autoinjecror
`
`Modeling of an autoinjection device involves a phys—
`ical understanding of the effects of all the components
`
`Vol. 65, No. 5, September—October 2011
`
`471
`
`REGITCOOO72198
`
`Regeneron Exhibit 1087.005
`
`

`

`Downloaded from journalpdaorg on May 31, 2012
`
`associated with the delivery system. The autoinjector
`system included in this study has an installed spring
`serving as the source of the driving force required to
`inject
`the product from the syringe. The spring is
`installed at a compressed length, which is shorter than
`its free length. At the time of injection, the spring is
`released from its installed length, causing the spring to
`relax while forcing the drug out of the syringe at the
`same time. The driving force from the spring at any
`time when the compressed spring length is “x” is given
`by
`
`Fspring : kiln " X)
`
`(9)
`
`where 10 is the free length of the spring, x is the current
`spring length, and k is the spring constant. A stronger
`spring will provide a higher driving force and a shorter
`injection time. The generalized equation (for both
`Newtonian and non—Newtonian fluids) for the momen—
`tum balance on the stopper can be written as
`
`
`dzx
`dx
`dx\ “
`Kh<
`)
`dt /
`stopper (“:2
`x) det
`
`“k(lo
`
`m
`
`(10)
`
`The terms Kf and Kh correspond to the frictional and
`the hydrodynamic terms, respectively, as described in
`
`the previous section. 1nSmpper is the mass of the stop-
`per, and n is the power law viscosity index of the
`liquid. Equation It) represents a one-dimensional, sec-
`ond—order differential equation capturing the motion
`of the stopper. it is based on the assumption that the
`fluid is always at
`the quasi—steady state where the
`hydrodynamic term corresponds
`to the Poiseuille
`equation.
`
`For Newtonian solutions (n = l), a reasonably accu—
`rate analytical solution for eq 10 can be obtained by
`using appropriate initial conditions (zero velocity for
`stopper) and applying assumptions including neglect—
`ing the inertia term and considering the system to be in
`a quasi-steady state where the spring force balances
`the hydrodynamic and friction forces. For Newtonian
`fluids, an analytical approximation can then be derived
`HS
`
`x(t) = l0 + (x0 — lo)exp
`
`n+mi
`
`rm
`
`in eq 10), another
`For non—Newtonian fluids (n + l
`assumption regarding the friction force can be made to
`
`derive an analytical solution. As reported later in
`Section D, friction force lies in the range of 1 to 3 N
`for a wide range of injection velocities (injection time
`of 3 to 30 s). For viscous products the hydrodynamic
`force term is significantly larger than the friction term
`and has a stronger dependence on injection speed. If
`the frictional term is assumed to be a constant (C r), an
`analytical solution for non—Newtonian fluids can be
`obtained and is given by
`
`
`
`x(t) =c0 + [—(n *1) k
`
`u/(u’l)
`
`(a~%wwfl
`
`on
`
`where
`
`Co :: lo "‘ Cf/k
`
`Results and Discussion
`
`Based on the theoretical framework presented in the
`previous sections, experiments were conducted to
`measure the parameters contributing to delivery forces
`and injection times, including syringe barrel diameter,
`needle diameter, and autoinjector spring constant. The
`friction forces were also estimated, along with the
`impact of stopper variability and injection speed. Once
`these parameters had been measured, delivery forces
`as estimated by eq 5 were verified with the experi—
`mental data generated using the lnstron. Injection time
`data for test autoinjectors were also compared to the
`calculated injection times as given by eq 11. Once the
`analytical model was confirmed to adequately capture
`the flow behavior inside an injection device, a theo—
`retical stack tolerance analysis was conducted to esti—
`mate the worst—case variability in injection time. The
`following sections describe the results from each of
`these assessments.
`
`A. Characterization of Barrel Internal Diameter
`
`Consistency in barrel size is important in estimating
`the delivery forces as stated in eq 5. The barrel size
`governs the area over which the force is applied to
`push the plunger rod during injection. The force has a
`fourth—power dependence on barrel radius for a given
`plunger speed (second—order dependence for a given
`flow rate). Table II shows the measured values for
`barrel diameters of different syringes. The measure-
`ments were conducted at different sections of syringes
`
`472
`
`PDA Journal of Pharmaceutical Science and Technology
`
`REGITCOOO72199
`
`Regeneron Exhibit 1087.006
`
`

`

`Downloaded from journalpdaorg on May 31, 2012
`
`TABLE II
`
`Vendor 2 (siliconized glass)
`
`
`
`
`
`
`
`
`Measurement of Barrel ID at Different Sections of Syringes
`
`Barrel ID Measurement (mean I standard deviation) mm
`1
`
`Syringe Vendor and Type
`Lot
`—l— Top (flange end) 4
`Middle
`Bottom (needle end)
`
`Vendor 1 (siliconized glass)
`Lot 1
`6.38 I 0.021
`1— 6.39 I 0.018
`6.38 I 0.019
`
`Lot 2
`6.38 : 0.024
`6.39 I 0.023
`[—
`6.38 : 0.025
`
`Lot 3 +_
`6.38 I 0.013
`+ 6.38 I 0.014 +
`6.38 I 0.016
`
`Lot 1
`6.34 I 0.023
`6.35 I 0.023
`6.34 I 0.020
`
`Lot 2
`6.35 I 0023
`L
`6.35 I 0.020
`6.35 I 0.017
`
`Lot 3
`6.35 I 0.018
`6.35 I 0.021
`:—
`6.35 I 0.019
`
`Lot 1+ 6.34 : 0.029
`T 6.34 I 0.027
`6.34 : 0.026
`’1
`Lot 2 f
`6.36 : 0.021
`T
`6.36 : 0.021
`6.36 : 0.021
`_.
`Lot 1
`6.21 I 0.024
`6.30 I 0.014
`6.34 I 0.007
`
`Vendor 3 (siliconized glass)
`
`Vendor 4 (plastic)
`
`in order to check the uniformity of the diameter
`
`throughout the syringe length. The “Top” section re—
`
`fers to the flange end and the “Bottom” refers to the
`needle end. Each data point in the table is based on
`measurements performed on a sample size of 20 sy—
`ringes. Data suggest reasonable consistency in barrel
`diameters across the syringe types (within 6.25 to 6.45
`mm). It is also observed that while the cross—sectional
`diameter of glass syringes is uniform along the barrel
`length. plastic (Vendor 4) syringes exhibit a small
`increase in barrel radius at the needle end. Based on
`
`the lots and sample size considered in this evaluation,
`the maximum variation of the barrel diameter in glass
`syringes is estimated to be around 1%, which can
`result in a variation of up to 4% in the hydrodynamic
`component of the extrusion force (eq 4) for a given
`plunger speed.
`
`B. Characterization of Needle Internal Diameter
`
`Needle lD also plays an important role in determining
`the net hydrodynamic force. A small Change in needle
`size can cause a significant change in delivery forces
`or injection time. Equation 5 shows that the force is
`dependent on the fourth power of needle radius.
`In—
`stead of measuring the absolute internal diameter, an
`indirect approach was used to estimate the variability
`in the internal diameter of needles. The method uses a
`
`flow—based set up in which a fluid (water) is pumped at
`constant [low rate through the needle and the pressure
`drop is measured at a steady state. Assuming needle
`lengths are consistent, the variability in the pressure
`drop provides an estimate of the variability in needle
`ED as
`
`Vol. 65, No. 5, September—October 2011
`
`Pressure dropM(
`
`Flow rate‘
`
`r:
`
`/l
`
`Or, variability in pressure drop and extrusion force :
`4 X variability in needle diameter.
`
`in reality, some variation in measured pressure drop
`across needles would be expected due to the variabil—
`ity in the needle lengths as well. This method was used
`to estimate the variability in the needle ID for 27 G
`syringes from different vendors based on a sample size
`of 10 units per data point. Figure 3 shows a box plot
`for measured force (normalized) for different syringe
`
`2
`
`1.8!
`1
`O)
`3 1.65'
`o
`‘14
`‘
`-e 1.4:
`0N)
`'7'; 12:
`L.
`a
`1
`‘o
`: é- é
`
`/
`
`/
`
`t/
`
`+
`
`Outliers
`,/
`a
`\
`.\
`/
`
`
`
`\\
`
`l
`
`.
`m
`Q 1
`I l
`3
`.‘
`Vendor‘s
`;
`
`T
`I
`E
`"'
`
`T
`I
`
`I
`"'
`
`\‘
`
`a
`
`+
`
`Vendor3
`
`Vendor 1
`0'81
`0.6" ' " '
`
`Vendor 2
`
`Figure 3
`
`A plot showing the variability in injection force
`(pressure drop) for multiple lots of different sy-
`ringe types. The force is normalized with the mean
`value of force from Vendor 2 (lot 1) and is reflective
`of needle radius.
`
`473
`
`REGITCOOO72200
`
`Regeneron Exhibit 1087.007
`
`

`

`Downloaded from journalpdaorg on May 31, 2012
`
`50*
`
`‘
`
`k = 0.30313 i 8.018291 N/mm(2.2 kgi)
`1;: 0.36032 3% 0.040112 N/mnm kgf)
`k= 0.55612 1 0.037004 N/mm(4.2 kgfl"
`
`systems. The boxes in the plot are drawn from first to
`third quartile with the center line being the median.
`The length of the bars (whiskers) is equivalent to 2.7
`times the standard deviation of each data set, and as a
`
`rule of thumb any point lying outside the whiskers is
`considered to be a statistical outlier. The force is
`normalized with the mean value of force for the Ven—
`
`dor 2 syringe (lot 1). The plot shows that Vendors 3
`and 4 syringes have the smallest needle diameter, as
`they have forces 60% higher than other syringes. This
`can cause up to a 60% increase in injection time for an
`autoinjector with these syringes (assuming the hydro-
`dynamic term is dominant). On the other hand,
`sy—
`ringes from Vendors l and 2 have similar forces,
`implying that the needle internal diameters are consistent
`among these syringes. Another important observation
`from the figure is that Vendors l and 2 have outlier
`needles that have up to 30% higher forces. This can
`result in up to a 30% increase in injection time.
`Therefore it is important to evaluate the variability
`in needle ID during the design of an autoinjector
`system. The frequency of such outlier needles could
`be lot-dependent, and its accurate estimation would
`require a larger sample set than that used in this
`study.
`
`C. Spring Force Characterization for Autoinjector
`
`The autoinjector system included in this study has an
`installed spring to serve as a source of the driving
`force required to inject the product from the syringe.
`To efficiently model the autoinjeetor, it is important to
`study the force—extension measurements and estimate
`the corresponding variability. For this purpose, mea—
`surements were performed using springs of different
`stiffness with a sample sine of 10 springs. Figure 4
`shows the plot of force-extension measurements for
`springs of three different spring constants. The shaded
`region indicates the length over which the spring re—
`mains active during injection. The measurements were
`performed using both a static and dynamic test recipe.
`For the dynamic measurement, the spring was gradu—
`ally compressed from its free length and the load was
`recorded as a function of length during compression.
`The plot in Figure 4 shows that the force extension
`profile is within the linear range as expected for
`spring—driven motion. An alternative method (static)
`was also employed in which the Instron compresses
`the spring to the installed length for a few seconds
`prior to force measurement. The spring was then com-
`pressed to the active length, paused. and then the
`
`
`
`70
`
`90
`
`110
`
`7 §.
`130
`150
`
`' ~,,
`170
`
`g 40»
`3—1
`g 30
`
`o m
`
`20»
`
`10»
`
`0,-
`30
`
`.
`
`Figure 4
`
`length (mm)
`
`Plot showing the dynamic force-extension measure-
`ments for springs of different stiffness (indicated
`by different color). The shaded region is the length
`over which the spring remains active during the
`time of injection.
`
`spring load was recorded. As shown in Table III, the
`force values are lower for the static test. This could
`
`potentially be attributed to the fact that static measure-
`ments allow the spring to relax and adjust to compres—
`sion, resulting in lower force relative to the dynamic
`measurement (where the spring is continually com—
`pressed). Either approach can be used to characterize
`the spring strengths, but the worst—case variability should
`be taken into account to design a robust autoinjector.
`
`D. Friction Force
`
`Variability in the inner barrel surface can be estimated
`by measuring the friction forces in empty syringes.
`Friction force can be estimated as the total force
`
`required for moving the plunger inside an empty sy—
`ringe. Measurement of friction force was performed
`using the Instron at an injection velocity ranging from
`3 to 900 mm/min. This velocity range corresponds to
`an injection time of 10 min to 3 s. respectively. for a
`1 mL injection. The wide range of injection times was
`chosen to evaluate very slow plunger movement that
`could potentially occur towards the end of a spring-
`driven autoinjector. Table IV shows the mean, maxi—
`mum, and range of friction force of syringes (includ—
`ing all
`lots) at an injection time of 6 5 (injection
`velocity 304.8 min/min) based on a sample size of 10
`units per syringe lot. The standard deviations listed in
`the table amount to a relatively large percent error and
`
`474
`
`PDA Journal of Pharmaceutical Science and Technology
`
`REGITCOOO72201
`
`Regeneron Exhibit 1087.008
`
`

`

`Downloaded from journalpdaorg on May 31, 2012
`
`TABLE III
`
`Variability in Spring Load Measurements
`
`Heavy Load (N)—Installed Length
`Light Load (N)—Activated Length
`
`Strength
`Static
`Dynamic
`Static
`4
`Dynamic
`
`A
`21.06 i 0.63
`25.72 i 1.91
`10.67 i 0.26
`11.24 i 0.85
`
`Spring
`
`
`
`
`16.07 i 1.92
`16.32 i 0.30
`32.15 I 5.80
`28.11 i 0.67
`l—
`B
`
`C
`l
`37.02 i 1.27
`45.06 i 3.37
`l8.98 : 0.49
`18.03 i 0.79
`i
`
`are the result of both syringe—to—syringe variability as
`well as the experimental error in measuring forces of
`such low magnitude.
`
`The profiles of the friction force as a function of
`displacement for siliconized glass syringes (Vendor 2)
`are shown in Figure 5(a) using an injection speed of
`304.8 mm/min (injection time of 6 s). The profiles
`correspond to a sample size of 20 data points for two
`syringe lots. The profile has an initial peak in the force
`that corresponds to the break—loose force. The profiles
`of friction force show that there could be significant
`variation in the lubrication in the glass syringes. it
`should be noted that for analysis purposes the mean
`value of the friction force (mean of the profile) is more
`representative of average lubrication of the syringe
`barrel. The maximum value can be used in the esti—
`mation of worst—case scenarios as well as in the de-
`
`tection of failure points.
`
`Figure 5(b) shows the dependence of friction force on
`injection velocities for different syringes. Friction
`force increases linearly with injection velocity, which
`is in agreement with the theory (eq 2). it should be
`noted that friction forces in syringes are less than 1%
`of the load cell (500 N) used for lnstron measure—
`ments. This could as a result contribute to large rela—
`tive error in measured friction forces, especially at low
`
`TABLE IV
`
`velocities. The friction force is slightly higher for
`plastic syringes, especially at low speeds, and shows a
`weaker dependence on injection speed. This is due to
`the fact that there is no lubrication in plastic syringes
`and thus the frictional force follows solid—solid cohe—
`
`sive behavior. Figure 5(c) shows the snapshot of vari-
`ability in friction force of different lots of syringes at
`an injection time of 6 s (which corresponds to an
`injection velocity of 304.8 min/min). The figure shows
`that
`the variability in friction force is comparable
`within each lot of syringes. The practical range of
`injection time is 30 to 3 s, and the friction force for
`this range of injection time varies from 1 to 3 N for all
`the syringes under consideration. Such variability in
`friction forces may not have a significant impact on
`injection times for high-viscosity products, where the
`hydrodynamic component of eq 1 is dominant. How—
`ever, for low-viscosity products, the frictional forces
`could be the key determinant of injection time and
`hence the syringe barrels and stoppers should be thor—
`oughly characterized to estimate the worst~case fric—
`tion force. It should also be noted that once the sy-
`ringes are filled with product,
`friction force may
`change due to the interaction between product and
`barrel surface. Such product—specific interactions re-
`quire measurement of friction force in a wettcd sy—
`ringe (product~filled syringe) and will be discussed in
`a separate article.
`
`Calculated Mean. Maximum, and Range of the Friction Force at an Injection Time of 6 s
`
`"T
`Injection Time: 6 s
`
`
`yringe
`m...—
`S
`.
`lV'ean
`Max
`Range
`
`Source/Type
`f (N)
`Std (N)
`f (N)
`l
`Std (N)
`min (N)
`max (N)
`Vendor 1
`1.79
`0.39
`208
`l
`0.44
`1.02
`3.20
`lT
`—1
`
`Vendor 2
`1.93
`0.40
`2.34
`0.47
`0.97
`3.40
`
`
`
`
`
`
`Vendor 3
`Vendor 4
`
`1.96
`2.34
`
`L
`
`0.38
`0.25
`
`l"
`
`2.20
`3.12
`
`0.48
`0.42
`
`]
`
`1.23
`1.55
`
`1'
`
`2.99
`3.83
`
`Vol. 65, No. 5, September—October 2011
`
`475
`
`REGITCOOO72202
`
`Regeneron Exhibit 1087.009
`
`

`

`Downloaded from journalpdaorg on May 31, 2012
`
`E Stopper Vendor A
`Stopper Vendor B
`
`2.5
`
`2
`
`1 5
`
`l
`
`0.5
`
`A a
`
`Ci
`.9
`
`.E
`LHLl—t
`
`0
`
`
`,
`fl
`1'
`'.‘
`*
`Syringe Vendor l
`Syringe Vendor 2
`
`Figure

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