throbber
The AAPS Journal, Vol. 19, No. 5, September 2017 ( # 2017)
`DOI: 10.1208/s12248-017-0126-0
`
`Review Article
`Theme: Nanotechnology in Complex Drug Products: Learning from the Past, Preparing for the Future
`Guest Editors: Katherine Tyner, Sau (Larry) Lee, and Marc Wolfgang
`
`Physicochemical Characterization of Iron Carbohydrate Colloid Drug Products
`
`Peng Zou,1,2 Katherine Tyner,1 Andre Raw,1 and Sau Lee1
`
`Received 7 April 2017; accepted 13 July 2017; published online 31 July 2017
`
`Iron carbohydrate colloid drug products are intravenously administered to
`Abstract.
`patients with chronic kidney disease for the treatment of
`iron deficiency anemia.
`Physicochemical characterization of iron colloids is critical to establish pharmaceutical
`equivalence between an innovator iron colloid product and generic version. The purpose of
`this review is to summarize literature-reported techniques for physicochemical characteriza-
`tion of iron carbohydrate colloid drug products. The mechanisms, reported testing results,
`and common technical pitfalls for individual characterization test are discussed. A better
`understanding of the physicochemical characterization techniques will facilitate generic iron
`carbohydrate colloid product development, accelerate products to market, and ensure iron
`carbohydrate colloid product quality.
`KEY WORDS: ferumoxytol; iron colloid; iron dextran; iron sucrose; sodium ferric gluconate.
`
`1 Center for Drug Evaluation and Research, US Food and Drug
`Administration, 10903 New Hampshire Ave, Silver Spring, Mary-
`land 20993, USA.
`2 To whom correspondence should be addressed. (e-mail:
`peng.zou@fda.hhs.gov)
`
`Abbreviations: AAS Atomic absorption spectroscopy, AFM Atomic
`force microscopy, AUC Analytical ultracentrifugation, BDI
`Bleomycin-detectable iron, DLS Dynamic light scattering, DPP
`Differential pulse polarography, DSC Differential scanning calorim-
`etry, EDX Energy-dispersive X-ray, EMR Electron magnetic
`resonance, EPR Electron paramagnetic resonance, ESA Electroki-
`netic sonic amplitude, ESR Electron spin resonance, EXAFS
`Extended X-ray absorption fine structure, FT-IR Fourier transform
`infrared spectroscopy, GPC Gel permeation chromatography,
`HPLC High-performance liquid chromatography, ICP-MS Induc-
`tively coupled plasma mass spectrometry, MDA Malondialdehyde,
`Maap Apparent molecular weight, Mn Number-average molecular
`weight, Mw Weight-average molecular weight, MPS Mononuclear
`phagocyte system, NEXAFS Near-edge X-ray absorption fine
`structure, NMR Nuclear magnetic resonance, NPP Normal pulse
`polarography, NTA Nitrilotriacetate, NTBI Nontransferrin-bound
`iron, OGD Office of Generic Drugs, SLS Static light scattering,
`SQUID Superconducting quantum interference device, STEM
`Scanning transmission electron microscope, TBA Thiobarbituric
`acid, TBI Transferrin-bound iron, TEM Transmission electron
`microscopy, TEM/NBED Transmission electron microscopy/nano
`beam electron diffraction, TEM/SAED Transmission electron
`microscopy/selected area electron diffraction, TGA Thermal gravi-
`metric analysis, USP United States Pharmacopeia, UV/Vis Ultravi-
`olet-visible spectroscopy, VSM Vibrating sample magnetometer,
`XAS X-ray absorption spectroscopy, XANES X-ray absorption
`near-edge structure, XRD X-ray diffraction
`
`INTRODUCTION
`
`Iron colloid drug products are intravenously adminis-
`tered as iron replacement therapies for the treatment of iron
`deficiency anemia in patients with chronic kidney disease
`receiving hemodialysis and supplemental epoetin therapy (1).
`After intravenous administration,
`iron colloids enter the
`bloodstream and are processed by the mononuclear phago-
`cyte system (MPS) (2). Once internalized by macrophages in
`the liver, spleen, and bone marrow, iron colloids are delivered
`to lysosomes and iron ions are released from the colloids and
`become part of the intracellular iron pool which is available
`for uses in biological processes. If
`iron is not needed
`immediately, the cell stores it in the form of ferritin or
`hemosiderin. If iron is needed elsewhere in the body, MPS
`will release iron from the cell and the extracellular protein
`transferrin will bind iron ions to form transferrin-bound iron
`(TBI) and deliver iron to where they are needed. If the serum
`transferrin is saturated with iron, the released iron weakly
`bind to serum components such as albumin to form
`nontransferrin-bound iron (NTBI) which causes oxidative
`stress and toxicity when taken up by the liver and heart (3).
`Therefore, the MPS uptake, in vivo stability, and iron release
`rate of
`iron colloids can dramatically impact
`their
`biodistribution, efficacy, and toxicity.
`Iron colloids are composed of an iron oxyhydroxide or
`iron oxide core and a complex carbohydrate coat with an
`average particle size ranging from 8 to 24 nm (4). The
`carbohydrate shell stabilizes and protects the iron core from
`hydrolysis, precipitation, and polymerization until the deliv-
`ery of iron colloids to the MPS. The carbohydrate shell also
`reduces the direct release of iron into the bloodstream before
`
`1359
`
`1550-7416/17/0500-1359/0 # 2017 American Association of Pharmaceutical Scientists
`
`PGR2020-00009
`Pharmacosmos A/S v. American Regent, Inc.
`Petitioner Ex. 1092 - Page 1
`
`

`

`1360
`
`NDA
`
`021135
`017807
`010787
`040024
`017441
`
`020955
`
`022180
`203565
`
`Table I. Parental Iron Colloid Product NDAs Approved by the US FDA
`
`Zou et al.
`
`Trade name
`
`Venofer®
`Proferdex®
`Imferon®
`Dexferrum
`INFeD®
`(Cosmofer® outside
`of the USA)
`Ferrlecit®
`
`Feraheme®
`Injectafer®
`(Ferinject® outside of
`North America)
`
`Generic name
`
`Iron sucrose
`Iron dextran
`Iron dextran
`Iron dextran
`Iron dextran
`
`Current sponsor
`
`Approved date
`
`Luitpold
`New River
`Fisons
`Luitpold
`Watson Labs
`
`November 06, 2000
`March 26, 1981
`April 25, 1957
`February 23, 1996
`April 29, 1974
`
`Sodium ferric gluconate
`complex in sucrose
`Ferumoxytol
`Ferric carboxymaltose
`
`Sanofi Aventis
`
`February 18, 1999
`
`AMAG Pharmas Inc.
`Luitpold in the USA and
`Vifor Pharma outside
`of North America
`
`June 30, 2009
`July 25, 2013
`
`NDA new drug application
`
`the iron colloids are delivered to the MPS and, hence,
`prevents the formation of NTBI and oxidative toxicity. As a
`result, the iron colloid products coated with various carbohy-
`in vivo
`drate shells exhibit differences in MPS uptake,
`stability, iron release profiles, and the content of labile and
`low molecular weight iron species (4).
`The intravenously administered iron colloid drug prod-
`ucts approved by the US FDA for the treatment of iron
`deficiency anemia are summarized in Table I. Differences
`exist among the formulations regarding the size of carbohy-
`drate group and iron colloids, surface property, stability of the
`iron-carbohydrate complex, and the rate at which iron is
`released from iron colloids. Among these approved iron
`colloid products,
`iron sucrose injection and iron dextran
`injection have individual United States Pharmacopeia (USP)
`monographs.
`The FDA has published product-specific bioequivalence
`draft guidances for the generic version of
`iron sucrose
`injection (5), ferumoxytol
`injection (6) and sodium ferric
`gluconate in sucrose injection (7), ferric carboxymaltose
`injection (8), and iron dextran injection (9). In the draft
`guidances, the applicants are recommended to conduct a
`comparative clinical bioequivalence study between the inno-
`vator product and generic version using total iron in serum
`and transferrin-bound iron in serum as pharmacokinetic
`endpoints. In addition, physicochemical characterization
`studies are recommended to demonstrate the same physico-
`chemical properties of iron colloids between the innovator
`product and generic version. The physicochemical properties
`which may affect the safety and efficacy of iron colloid drug
`products are summarized in Table II and divided into three
`categories:
`the whole nanoparticle properties,
`iron core
`properties, and carbohydrate shell properties. These physico-
`chemical properties are likely correlated with the pharmaco-
`kinetics and tissue distribution of iron colloids and may also
`impact in vivo stability and iron release kinetics.
`The European Medicines Agency (EMA) expresses
`different opinions on the data required to demonstrate
`bioequivalence of intravenous iron colloid products (10). In
`addition to comparative physicochemical characterization and
`plasma pharmacokinetic study in humans, EMA recommends
`
`nonclinical studies to compare tissue distribution and toxicity
`of generic and innovator iron colloid drugs in animal models.
`The tissue distribution study includes the assessment of iron
`distribution in at least three biological compartments: plasma
`and red blood cells, reticulo-endothelial system (RES) (liver
`and spleen), and pharmacological/toxicological target tissues
`(bone marrow, kidney, liver, lung, or heart) in animal models.
`In case that minor differences are observed in physicochem-
`ical characterization, nonclinical or human pharmacokinetic
`studies, a therapeutic equivalence study is recommended to
`address their impact on efficacy and safety.
`Although the critical physicochemical properties have
`been recommended in the FDA product-specific draft guid-
`ances, the specific techniques for iron colloid characterization
`have not been systemically reviewed. The availability of
`reliable characterization techniques is critical for demonstrat-
`ing the same physicochemical properties of two iron colloid
`products. To fill the knowledge gap on iron colloid charac-
`terization, we have summarized the physicochemical tests that
`have been reported in the literature to characterize iron
`colloid drug products (Table II). This review provides an
`overview of currently available iron colloid characterization
`techniques and a summary of recent advances in physico-
`chemical characterization of iron colloids, which may facili-
`tate the development of iron colloid drug products and assure
`drug product quality.
`
`CHARACTERIZATION OF PHYSICOCHEMICAL
`PROPERTIES OF THE WHOLE PARTICLES
`
`After an iron colloid enters systemic circulation via
`intravenous infusion,
`the particle is tagged by plasma
`proteins for recognition (opsonization) and processed by
`the MPS system. After
`internalized by marcrophages
`residing in the liver and spleen,
`iron colloids accumulate
`in lysosomes and release free irons. The tissue distribution,
`iron release kinetics, efficacy, and safety of iron colloids are
`affected by a number of physicochemical properties of the
`whole particles including chemical composition of
`iron
`colloids, molecular weight distribution, particle size and
`distribution, content of
`low molecular weight
`iron and
`
`PGR2020-00009
`Pharmacosmos A/S v. American Regent, Inc.
`Petitioner Ex. 1092 - Page 2
`
`

`

`Physicochemical Characterization of Iron Colloids
`
`1361
`
`Table II. Physicochemical Characterization of Iron Complex Drugs
`
`Attributes
`
`Reasons to recommend
`the attributes
`
`Tests
`
`Whole particle
`
`Equivalence in stoichiometric
`ratios of iron, free and
`bound carbohydrate, and
`other relevant components
`Molecular weight distribution
`(Mw, Mn, and Mw/Mn)
`Low molecular weight iron species
`
`Labile iron
`
`Particle size distribution
`
`Iron core
`
`Iron core size and morphology
`
`Crystallinity
`
`Iron environment (valence state
`of iron, the spin state of ferric,
`and the type of coordination state
`of iron atoms and details
`of ligand binding)
`Fe3+ to Fe2+ reduction potential
`and Fe(II) content
`Magnetic properties
`
`Carbohydrate shell Carbohydrate composition and
`carbohydrate-iron
`core interactions
`
`Surface charge
`
`Characterization of
`polysaccharides
`
`Q1/Q2
`
`USP requirement
`
`Formation of NTBI
`and potential oxidative toxicity
`Formation of NTBI and
`potential oxidative toxicity
`Related to opsonization, MPS
`uptake, PK and tissue distribution
`Related to iron release rate, MPS
`uptake, PK and tissue distribution
`Related to iron release
`and in vivo stability
`Related to iron release
`and in vivo stability
`
`Iron assay (atomic absorption),
`carbohydrate assay (HPLC), or
`elemental analysis
`(AAS, ICP-MS, or EDX)
`AUC or GPC
`
`Dialysis or ultrafiltration
`
`Bleomycin assays or iron
`chelation assays
`DLS or AFM
`
`TEM (diameter), AFM, Mössbauer
`spectroscopy, and XRD
`Mössbauer spectroscopy, Raman, XRD,
`TEM/SAED or TEM/NBED, XANES
`Mössbauer spectroscopy, EPR,
`Raman, and UV/Vis
`
`Related to in vivo stability and to
`detect impurities
`To measure overall sameness
`in iron core and detect impurities
`Related to opsonization,
`MPS uptake, PK,
`tissue distribution, iron release
`rate and in vivo stability
`Related to opsonization,
`MPS uptake, PK, and
`tissue distribution
`Related to in vivo stability
`
`Polarography, cerimetric titration
`
`VSM or SQUID
`
`FT-IR, thermal analysis
`(TGA or DSC), changes in
`particle size under dilution,
`or polarography
`Electrophoretic mobility
`(e.g., zeta potential),
`potentiometric titration
`NMR, size exclusion
`chromatography, copper assay
`
`HPLC high-performance liquid chromatography, AAS atomic absorption spectroscopy, ICP-MS inductively coupled plasma mass
`spectrometry, EDX energy-dispersive X-ray, Mn number-average molecular weight, Mw weight-average molecular weight, USP United States
`Pharmacopeia, AUC analytical ultracentrifugation, GPC gel permeation chromatography, NTBI nontransferrin-bound iron, MPS mononuclear
`phagocyte system, DLS dynamic light scattering, AFM atomic force microscopy, TEM transmission electron microscopy, XRD X-ray
`diffraction, TEM/SAED transmission electron microscopy/selected area electron diffraction, TEM/NBED transmission electron microscopy/
`nano beam electron diffraction, XANES X-ray absorption near-edge structure, EPR electron paramagnetic resonance, VSM vibrating sample
`magnetometer, SQUID superconducting quantum interference device, FT-IR Fourier transform infrared spectroscopy, TGA thermal
`gravimetric analysis, DSC differential scanning calorimetry, NMR nuclear magnetic resonance
`
`labile iron, and stability of iron colloids. The characteriza-
`tion of
`these physicochemical properties of
`the whole
`particles is discussed in this section.
`
`Stoichiometric Ratios of Iron, Free and Bound Carbohydrate,
`and Other Excipients
`
`The iron in iron colloids is measured by ICP-MS, atomic
`absorption spectroscopy, or atomic emission spectroscopy,
`while the unbound carbohydrates are measured by chromato-
`graphic methods. The presence of unbound and loosely
`bound carbohydrates,
`low molecular weight Fe complex,
`excipients, and impurities may mask the real elemental
`composition in iron colloids. It is desired to completely
`remove these interfering species by dialysis or other
`
`approaches prior to elemental analysis. The amounts of
`carbon and hydrogen in dialyzed iron colloids can be
`measured by CHN analysis, energy dispersive X-ray elemen-
`tal analysis, atomic absorption, or other elemental analysis
`techniques.
`Elemental analysis of dialyzed iron core obtained from
`Venofer® and Ferrlecit® has been reported by Kudasheva
`et al. (11). From the molar ratio of iron to carbon (1.45:1) for
`Venofer® determined by elemental analysis, they concluded
`that one sucrose molecule of 12 carbon atoms is present for
`17 atoms of iron. For iron-gluconate complex, the elemental
`analysis gives a molar ratio of iron to carbon as 1.30:1,
`indicating that one gluconate of 6 carbon atoms is present for
`7.8 atoms of Fe. Considering the molecular formula of
`Ferrlecit® [NaFe2O3(C6H11O7)(C12H22O11)5]n = 200,
`the
`
`PGR2020-00009
`Pharmacosmos A/S v. American Regent, Inc.
`Petitioner Ex. 1092 - Page 3
`
`

`

`1362
`
`Zou et al.
`
`measured gluconate to iron ratio (1:7.8) revealed that only a
`small fraction of total gluconate binds to the iron core.
`Accurate elemental analysis depends on a complete removal
`of unbound and loosely bound carbohydrates, low molecular
`weight Fe complex, excipients, and impurities. In another
`reported study (12), the percentages (wt%) of C, H, N, and
`Fe in dialyzed Venofer® were determined as 36.31 ± 0.33,
`5.15 ± 1.87, 0.08 ± 0.12, and 5.7 ± 0.3, respectively, indicating
`that one sucrose molecule of 12 carbon atoms is present for
`2.5 atoms of iron. The high percentage of carbon indicated
`that the sample dialysis was likely incomplete, illustrating
`some of
`the limitations of elemental analysis for the
`characterization of iron colloids.
`
`Molecular Weight Distribution
`
`The average molecular weight and molecular weight
`distribution are indicators of the quality and stability of iron
`colloids and can be used to detect variations in product
`quality and potential degradation or aggregation of iron
`colloids. Usually,
`the weight-average molecular weight
`(Mw), the number-average molecular weight (Mn), and the
`polydispersity index (Mw/Mn) are measured to reflect the
`molecular weight distribution of the iron colloids. In the
`USP monograph of iron sucrose injection, the acceptance
`criteria of molecular weight are as follows: Mw = 34,000–
`60,000 Da, Mn ≥ 24,000 Da, and Mw/Mn ≤ 1.7. The
`apparent molecular weight of sodium ferric gluconate in
`sucrose complex (Maap 289,000–440,000 Da) is provided in
`the label of Ferrlecit®. In addition to Mw, Mn, and Mw/Mn,
`the Maap of sodium ferric gluconate in sucrose complex
`needs to be measured (13,14). It is inappropriate to directly
`compare the molecular weight of iron sucrose and sodium
`ferric gluconate in sucrose complex because of the differ-
`ences in analytical methods (chromatographic conditions,
`columns, and calibration standards). The common analytical
`methods for determining molecular weight distribution of
`iron colloids include size exclusion chromatography (SEC)
`or gel permeation chromatography (GPC), analytical ultra-
`centrifugation (AUC), and static light scattering (SLS). For
`these analytical methods,
`the selection of appropriate
`molecular weight calibration standards and calibration
`curve plotting method are crucial for accurate measurement
`of iron colloid molecular weight.
`
`Particle Size and Size Distribution
`
`Among the techniques used to determine the particle
`size distribution of carbohydrate-coated iron colloids, dy-
`namic light scattering (DLS) is most widely used. The
`particle size of iron colloids measured by DLS is summa-
`rized in Table III (4). DLS measurement is quick and
`sample preparation is very simple. However, the results are
`affected by the experimental conditions. For comparative
`particle size measurements, samples should be measured
`under the same experimental conditions. The DLS method
`should be appropriately validated for robustness and
`accuracy. The existence of free sucrose and/or gluconic acid
`may probably affect the particle size measured by DLS (11).
`
`is desired to measure the size and size
`it
`Therefore,
`distribution of samples by DLS before and after sample
`dialysis. It was reported that the removal of free carbohy-
`drates did not cause aggregation or degradation of iron
`colloids (11,14,18).
`
`Low Molecular Weight Iron
`
`Low molecular weight iron (or Bfree iron^) is the iron
`impurities which can be physically separated from iron
`colloids by dialysis or ultrafiltration (4,13,15,19). One of
`the concerns for iron colloid drug products is the potential
`toxicity of low molecular weight iron in the formulations
`(20). According to the USP monograph for iron sucrose
`injection (21), no additional peaks in the polarogram can
`be used to demonstrate the absence of
`low molecular
`weight iron complex. However, the quantitation of
`low
`molecular weight
`iron species by polarography is not
`sensitive (22). Low molecular weight
`iron can also be
`measured by dialysis or ultrafiltration. In the label of
`sample dialysis against a 12–14-kDa cutoff
`Ferrlecit®,
`membrane over a period of up to 270 min is recom-
`mended for the measurement of
`low molecular weight
`iron and the acceptable limit is 1% of total iron. Previous
`studies (Table IV) have revealed that the amount of low
`m o l e c u l a r w e i g h t
`i r o n
`r a n k s
`a s
`f o l l o w s :
`Ferrlecit® > Venofer® ≈ INFeD® > Feraheme®,
`Injectafer®, and Monofer® (4,13).
`The amount of low molecular weight iron determined
`by ultrafiltration depends on filter cutoff, pH of the buffer,
`and centrifugation speed. The amount of low molecular iron
`determined by dialysis depends on the components in
`dialysis buffer, the buffer pH, the volume ratio of iron
`colloid suspension and dialysis buffer, and the cutoff of
`dialysis membrane. For example, a higher percentage of
`dialyzable iron was detected when 0.9% sodium chloride
`solution at pH 7.5 was used as the dialysis buffer compared
`with pH 7.5 water (4). Therefore, a comparative quantita-
`tion of low molecular weight iron complex in two iron
`colloid samples should be conducted under the same
`experimental conditions and in the presence of control
`samples.
`
`Labile Iron
`
`Different from dialyzable low molecule weight iron or free
`iron, labile iron is iron impurities which physically bind to iron
`carbohydrate matrix and are not dialyzable in saline but readily
`released in blood circulation (4,15). In the presence of 4.5%
`bovine serum albumin in dialysis buffer (Table IV), however,
`the amount of dialyzable iron dramatically increased (13),
`indicating that labile iron can be dialyzed in the presence of
`plasma protein or iron chelation agents (4). If serum transferrin
`is saturated, the released labile iron in the bloodstream forms
`NTBI, which potentiates oxidative stress and inflammation, then
`resulting in direct cellular damage and possibly increasing the
`risk of atherosclerotic disease (24). The content of labile iron
`can be measured under physiological conditions using several
`reported assays, which are divided into two categories:
`bleomycin assays and chelation-based assays (25). As shown in
`Table IV, the content of labile iron roughly follows the sequence:
`
`PGR2020-00009
`Pharmacosmos A/S v. American Regent, Inc.
`Petitioner Ex. 1092 - Page 4
`
`

`

`Physicochemical Characterization of Iron Colloids
`
`1363
`
`Table III. The Mean Size of Iron Core and Whole Particle of Iron Colloids
`
`Iron colloid products
`
`Mean diameter of iron core (nm)
`
`Mean diameter of whole particle (nm)
`
`Ferrlecit
`Venofer
`Feraheme
`INFeD
`Injectafer
`Monofer
`Refs.
`
`TEM
`
`4.1 ± 1.7
`5.0 ± 0.8
`6.2 ± 1.4
`5.6 ± 1.2
`11.7 ± 4.4*
`6.3 ± 1.2
`(4)
`
`XRD
`
`Mössbauer
`
`3.4
`3.3
`6.4
`4.4
`4.3
`4.2
`(4)
`
`N.A.
`2.5
`5–10
`N.A.
`3.3
`2.5
`(15,16)
`
`AFM
`
`3 ± 1
`4 ± 2
`N.A.
`N.A.
`N.A.
`N.A.
`(11,17)
`
`DLS
`
`8.6
`8.3
`23.6
`12.2
`23.1
`9.1
`(4)
`
`* The mean diameter of an agglomeration of several cores. Single cores are not definable. TEM transmission electron microscopy, XRD X-ray
`diffraction, AFM atomic force microscopy, DLS dynamic light scattering, N.A. not available
`
`Ferrlecit®≈Venofer®>INFeD®>Dexferrum®>Feraheme®≈Inj-
`ectafer® ≈ Monofer®. The content of labile iron measured by
`bleomycin assay is lower than that determined by EDTA
`chelation assay or transferrin binding assay (Table IV). This
`might probably be explained that the bleomycin assay specifi-
`cally detects redox-active iron only while the chelation-based
`assay detects the total amount of labile iron and free iron (25).
`Both bleomycin assay and chelation assay usually cannot
`distinguish free iron and labile iron, and the amount of free
`iron is included in the measured labile iron (15). The values of
`labile iron assays are dependent on binding affinity of selected
`chelators and the concentration of chelators. Therefore, large
`cross-lab variability in labile iron measurement is usually
`observed. It is important to compare the labile iron of two iron
`colloid products using the same assay and under the same
`experimental conditions.
`
`Bleomycin Assay
`
`The conventional bleomycin assay was first reported for
`the selective measurement of radical-promoting,
`loosely
`bound iron in biological fluids (26), and the detected iron is
`generally named bleomycin-detectable iron (BDI).
`Bleomycin is an anticancer drug which induces DNA
`degradation in the presence of ferrous ion. In a conventional
`bleomycin assay, bleomycin is incubated with DNA, ascorbic
`acid which converts ferric ions into ferrous ions, and the
`biological sample. The formation of bleomycin-Fe(II) com-
`plex causes degradation of DNA. DNA damage is then
`quantified by the formation of malondialdehyde (MDA) from
`the deoxyribose moiety of the DNA using the thiobarbituric
`acid (TBA) test. The formation of (TBA)2-MDA adduct
`causes UV absorption at 532 nm, and therefore,
`the
`
`Table IV. Contents of Low Molecular Weight Iron and Labile Iron in Iron Colloid Drug Products
`
`Iron colloid drugs
`
`Low molecular
`weight iron (free iron)
`
`Free iron + labile iron
`
`Dialysis in pH
`7.5 NaCl
`solution (4)
`
`Ultra-
`fi l t r a t i o n
`(13)
`
`Dialysis in 4.5%
`albumin (13)
`
`Bleomycin
`assay (13)
`
`Transferrin
`binding (100–125 mg
`dose) (19)
`
`Transferrin
`binding (200 mg
`dose) (4)
`
`EDTA
`c h e l a t i o n
`(23)
`
`Iron dextran
`(Dexferrum®
`or DexIron®)
`Low MW
`iron dextran
`(INFeD®)
`Sodium ferric
`gluconate
`(Ferrlecit®)
`Iron sucrose
`(Venofer®)
`Ferumoxytol
`(Feraheme®)
`Iron
`carboxymaltose
`(Injectafer®)
`Iron isomaltoside
`(Monofer®)
`
`–
`
`–
`
`–
`
`–
`
`2.5%
`
`0.207%
`
`0.30%
`
`0.70%
`
`0.19%
`
`3.4%
`
`1.338%
`
`2.36%
`
`4.80%
`
`1.40%
`
`5.8%
`
`0.067%
`
`<0.002%
`
`<0.002%
`
`<0.002%
`
`0.038%
`
`0.001%
`
`2.80%
`
`0.55%
`
`–
`
`–
`
`–
`
`–
`
`0.69%
`
`0.07%
`
`–
`
`–
`
`4.5%
`
`–
`
`–
`
`–
`
`–
`
`2.1%
`
`3.2%
`
`3.5%
`
`0.8%
`
`0.5%
`
`0.9%
`
`–
`
`0.8%
`
`3.6%
`
`3.8%
`
`–
`
`–
`
`–
`
`PGR2020-00009
`Pharmacosmos A/S v. American Regent, Inc.
`Petitioner Ex. 1092 - Page 5
`
`

`

`1364
`
`Zou et al.
`
`concentrations of BDI in the biological sample can be
`quantified using a standard curve of the UV absorption at
`532 nm (27). However, the formation of (TBA)2-MDA
`adduct requires relatively harsh conditions. For example, the
`mixture is incubated at 80°C for 20 min to allow the
`formation of (TBA)2-MDA adduct. Side reactions occur at
`high temperature, which may result in the generation of
`misleading data (27).
`To overcome the limitations of conventional bleomycin
`assay, Burkitt et al. modified the assay by using ethidium-
`binding to measure DNA damage caused by bleomycin-
`Fe(II) complex (27). Ethidium bromide is a dye for DNA
`staining. The intercalation of ethidium bromide into DNA
`enhances fluorescence, while the DNA damage caused by
`bleomycin-Fe(II) complex can decrease the fluorescence
`intensity. A nonlinear calibration curve is established between
`BDI concentration and fluorescence intensity. The ethidium-
`binding-based bleomycin assay was found particularly re-
`sponsive to damage induced by iron at very low concentration
`end of the standard curve, but saturation occurred at higher
`concentrations. The range of
`iron concentration in the
`standard curve can be adjusted by changing the concentration
`of DNA in the incubation.
`
`Iron Chelation Assays
`
`labile iron is first
`For iron chelation-based assays,
`mobilized with a chelating agent and then the chelation
`complex is separated (optional) and measured by different
`methods, including HPLC, colorimetric measurement, fluo-
`rescence measurement, and atomic absorption spectropho-
`tometry (AAS). The reported chelation agents for labile iron
`assay include Ferrozine® (4), EDTA (23), nitrilotriacetate
`(NTA) (25), oxalate (28), chromazurol B (29), bathophenan-
`throline (30), polyphenols in tea extract (15), and others.
`The most widely used chelation agent for labile iron
`determination is Ferrozine®, which can detect not only low
`molecular weight iron but also labile iron and transferrin-
`bound iron in serum (4,19). Briefly, the iron colloid sample
`is incubated in human plasma at room temperature, and
`then the mixture is added with detergent, thiourea, and
`citric acid followed by the addition of sodium ascorbate and
`Ferrozine. The detergent serves to clarify plasma sample,
`acidic buffer lowers pH to <2 to free Fe(III) from
`transferrin, and ascorbic acid reduces Fe(III) to Fe(II)
`which forms complex with Ferrozine. The UV absorption
`of the complex is measured at 562 nm. One concern for this
`procedure is that the harsh conditions such as low pH may
`cause degradation of iron colloids, resulting in overestima-
`tion of labile iron. To avoid the harsh conditions, Van Wyck
`et al. (19) removed iron colloids from the plasma incubation
`mixture using an alumina column before the Ferrozine test.
`Only the protein-bound labile iron and protein-bound free
`iron were measured.
`A capillary electrophoresis method coupled with EDTA
`chelation was used to determine labile iron in iron colloid
`samples (23). Different from Ferrozine which formed Fe(II)-
`chelate, EDTA formed Fe(III)-EDTA chelate and therefore
`reduction with ascorbic acid was not required. The Fe(III)-
`EDTA chelate was separated on a capillary electrophoresis
`and detected at 246 nm.
`
`For the assay of labile iron in iron sucrose and sodium
`ferric gluconate complex in sucrose, acidic conditions can
`cause hydrolysis of sucrose to form glucose and fructose. Due
`to their free ketone group, glucose and fructose can reduce
`Fe(III) in the assay to Fe(II). Therefore, when a Fe(III)
`chelate is used to measure labile iron of iron sucrose and
`sodium ferric gluconate complex in sucrose samples under
`acidic conditions, the autoreduction by reducing sugars may
`result in underestimation of labile iron.
`
`IRON CORE CHARACTERIZATION
`
`The efficacy and safety of iron colloids are affected by
`not only the properties of the whole particle but also the iron
`core properties. For example, macrophage uptake and tissue
`distribution of iron colloids may be impacted by the iron core
`size and morphology (31). The in vivo stability and iron
`release rate are correlated with iron core size, crystallite
`structure, and iron core environment. For iron colloids with
`magnetic property, comparative magnetic characterization
`can detect the potential impurities and subtle differences in
`iron core structure and environment (32).
`
`Iron Core Size and Morphology
`
`Iron carbohydrate complexes are composed of an iron
`core surrounded by associated carbohydrate ligands. Al-
`though DLS can be used to determine the hydrodynamic size
`of iron colloids, it is unable to measure the size of the iron
`core. The diameter of the iron core can be measured by
`transmission electron microscopy (TEM), X-ray diffraction
`(XRD), Mössbauer spectroscopy, and other appropriate
`techniques. TEM and atomic force microscopy (AFM) are
`frequently used to investigate the particle morphology (e.g.,
`shape and agglomeration status of the particles).
`
`Transmission Electron Microscopy
`
`TEM is frequently used to measure the diameter and
`morphology of the iron core of iron colloid products. The
`carbohydrate shell usually cannot be detected by TEM
`without appropriate staining due to the lower electron density
`of the carbohydrate shell. Jahn et al. reported the mean
`diameter of Venofer®, Ferrlecit®, Feraheme®, INFeD®,
`Injectafer®, and Monofer® iron core as 5.0 ± 0.8, 4.1 ± 1.7,
`6.2 ± 1.4, 5.6 ± 1.2, 11.7 ± 44, and 6.3 ± 1.2 nm, respectively
`(Table III) (4). It is worthy to note that the mean diameter of
`the Injectafer® particle is the mean diameter of an agglom-
`eration of several
`iron cores. Individual
`iron cores of
`Injectafer are not able to be resolved under TEM. The
`average iron core size of Feraheme® measured by TEM was
`reported by Balakrishnan et al. as 6.4 ± 0.4 nm (13). Although
`it was difficult to accurately determine the iron core sizes of
`Venofer®, Ferrlecit®, and INFeD® due to an undefined iron
`core margin, Balakrishnan et al. ranked the average core size
`as follows: Feraheme® > INFeD® > Venofer® ≥ Ferrlecit®
`(13). Inconsistent with Jahn’s results, Kudasheva et al. (11)
`reported the iron core size distribution of dialyzed Venofer
`and Ferrlecit ranging from 1.0 to 6.5 nm (mean 3 ± 2 nm) and
`0.9 to 3.5 nm (mean 2 ± 1 nm), respectively. Bullivant et al.
`reported the mean diameter of Feraheme® core as 3.25 nm
`
`PGR2020-00009
`Pharmacosmos A/S v. American Regent, Inc.
`Petitioner Ex. 1092 - Page 6
`
`

`

`Physicochemical Characterization of Iron Colloids
`
`1365
`
`Fig. 1. Cryo-TEM micrographs of a Ferrlecit®, b Venofer®, c INFeD®, and d Feraheme® (34)
`
`(33). These inconsistencies are likely caused by the differ-
`ences in sample concentrations, sample preparation, instru-
`ment variability, instrumental parameters, and data analysis.
`Morphology analysis is often performed concurrently
`with the iron core size determination. A recent study
`conducted at an FDA lab showed that cryo-TEM where the
`samples were prepared in liquid nitrogen could generate a
`higher resolution image (Fig. 1) and a more accurate
`measurement of iron core size compared with room temper-
`ature TEM because the sample dehydration at room temper-
`ature may potentially cause aggregation (34). The cryo-TEM
`data showed that the iron cores of Venofer®, Ferrlecit®,
`INFeD®, and Feraheme® share a spherical morphology. As
`shown in Fig. 1d, Feraheme® exhibits clusters of multiple
`particles rather than individual particles. The clusters were
`observed in the cryo-TEM images of both undiluted
`Feraheme® sample and 100-fold diluted sample. Consistently,
`clusters of multiple particles were also observed in TEM and
`STEM images of Feraheme® a

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