throbber
Journal of Inorganic Biochemistry 98 (2004) 1757–1769
`
`JOURNAL OF
`
`Inorganic
`Biochemistry
`
`www.elsevier.com/locate/jinorgbio
`
`Structure of carbohydrate-bound polynuclear iron oxyhydroxide
`nanoparticles in parenteral formulations
`
`Dina S. Kudasheva a,b, Jriuan Lai a,b,c, Abraham Ulman a,b,c, Mary K. Cowman a,b,*
`
`a Othmer Department of Chemical and Biological Sciences and Engineering, Polytechnic University, Six Metrotech Center, Brooklyn, NY 11201, USA
`b Herman F. Mark Polymer Research Institute, Polytechnic University, Six Metrotech Center, Brooklyn, NY 11201, USA
`c The NSF MRSEC for Polymers at Engineered Interfaces, Polytechnic University, Six Metrotech Center, Brooklyn, NY 11201, USA
`
`Received 21 January 2004; received in revised form 28 April 2004; accepted 11 June 2004
`Available online 11 September 2004
`
`Abstract
`
`Intravenous iron therapy is used to treat anemia associated with chronic kidney disease. The chemical structures of parenteral
`iron agents have not been characterized in detail, and correlations between structure, efficiency of iron delivery, and toxicity via
`catalysis of oxygen-derived free radical creation remain to be established. In this study, two formulations of parenteral iron have
`been characterized by absorption spectroscopy, X-ray diffraction analysis (XRD), transmission electron microscopy (TEM), atomic
`force microscopy (AFM), and elemental analysis. The samples studied were VenoferÒ (Iron Sucrose Injection, USP) and FerrlecitÒ
`(Sodium Ferric Gluconate in Sucrose Injection). The 250–800-nm absorption spectra and the XRD patterns showed that both for-
`mulations contain a mineral core composed of iron oxyhydroxide in the b-FeOOH mineral polymorph known as akaganeite. This
`was further confirmed for each formulation by imaging using TEM and AFM. The average core size for the nanoparticles, after
`dialysis to remove unbound or loosely bound carbohydrate, was approximately 3 ± 2 nm for the iron–sucrose, and approximately
`2 ± 1 nm for the iron–gluconate. Each of the nanoparticles consists of a mineral core, surrounded by a layer of bound carbohydrate.
`The overall diameter of the average bead in the dialyzed preparations was approximately 7 ± 4 nm for the iron–sucrose, and 3 ± 1
`nm for the iron–gluconate. Undialyzed preparations have particles with larger average sizes, depending on the extent of dilution of
`unbound and loosely bound carbohydrate. At a dilution corresponding to a final Fe concentration of 5 mg/mL, the average particle
`diameter in the iron–sucrose formulation was approximately 22 ± 9 nm, whereas that of the iron–gluconate formulation was
`approximately 12 ± 5 nm.
`Ó 2004 Elsevier Inc. All rights reserved.
`
`Keywords: Parenteral formulations; Iron oxyhydroxide; Iron–carbohydrate complex; AFM; Iron–sucrose injection; Sodium ferric gluconate;
`Akaganeite
`
`1. Introduction
`
`Parenteral iron formulations containing iron com-
`plexed to monosaccharides or disaccharides are among
`the products currently used for the treatment of anemia.
`
`* Corresponding author. Tel.: +1 718 260 3054; fax: +1 718 260
`3125.
`E-mail address: mcowman@poly.edu (M.K. Cowman).
`
`0162-0134/$ - see front matter Ó 2004 Elsevier Inc. All rights reserved.
`doi:10.1016/j.jinorgbio.2004.06.010
`
`Yet, the chemical structures of these materials have not
`been characterized in detail, and correlations between
`structure and complex stability, efficiency of iron deliv-
`ery, and toxicity via catalysis of oxygen-derived free rad-
`ical creation remain to be established. This study
`represents a first step toward complete characterization
`of the structures of this subclass of modern therapeutic
`iron–carbohydrate compounds.
`Iron(III) oxyhydroxide is formed in aqueous solu-
`tions by the dissolution of a suitable iron salt, such as
`ferric chloride, at low pH, followed by neutralization
`
`Pharmacosmos A/S v. Luitpold Ex. Pharmaceuticals, Inc., IPR2015-01490
`
`Luitpold Pharmaceuticals, Inc., Ex. 2048, P. 1
`
`

`
`1758
`
`D.S. Kudasheva et al. / Journal of Inorganic Biochemistry 98 (2004) 1757–1769
`
`or basification [1–3]. The aquated iron ions are hydro-
`lyzed initially to the mononuclear iron hydroxides
`Fe(OH)(H2O)5 and Fe(OH)2(H2O)4. Formation of
`Fe(OH)3(H2O)3 leads to the growth of a polynuclear
`(containing multiple linked iron atoms) iron oxyhydrox-
`ide with the approximate composition FeOOH. This
`mineral is insoluble unless stabilized by the presence of
`a water-soluble chelating agent.
`Carbohydrates can be appropriate chelating agents for
`the stabilization of iron oxyhydroxide nanoparticles in
`colloidal suspension. Certain low molecular mass neutral
`carbohydrates, such as sucrose [4], or fructose [5] present
`multiple hydroxyl groups in a suitable array to chelate
`iron, although the binding is inherently weak in neutral
`aqueous solution. Polymeric carbohydrates such as dex-
`tran may also be used to stabilize nanoparticles because
`they present a large number of hydroxyl groups, which
`can cooperatively chelate the surface of iron oxyhydrox-
`ide nanoparticles. For a neutral carbohydrate, the chela-
`tion to iron is enhanced at high pH, because the hydroxyl
`groups may become deprotonated, thus acquiring a neg-
`ative charge, and interacting more strongly with the cati-
`onic iron ion [6]. At neutral pH,
`inherently anionic
`carbohydrates such as gluconate or partially oxidized
`polysaccharides are more effective nanoparticle stabiliz-
`ers. In these cases, a carboxyl group provides the negative
`charge over a broad pH range. Its chelation to iron can
`also drive the deprotonation of nearby hydroxyl groups,
`further enhancing the complex stability [4,7].
`The structures of complexes formed between iron
`nanoparticles and polysaccharides have been studied
`previously, because iron–dextran complexes [8,9] and
`iron–polymaltosate complexes [10–12] (where polymal-
`tosate was the commercial name for partially hydro-
`lysed and oxidized starch) have been used for
`decades in the treatment of anemia. In a number of
`cases, the nanoparticle cores have been identified as
`iron oxyhydroxide. X-ray diffraction data are most
`consistent with the akaganeite (b-FeOOH) polymorph
`(polymorphs are solids with the same chemical com-
`position but different crystal structures) [13–18]. Ultra-
`violet–visible
`(UV–VIS)
`absorption
`spectroscopy
`[5,17–19] and Mo¨ ssbauer spectroscopy [13–16,20] have
`shown the presence of octahedrally coordinated high-
`spin Fe(III)
`ions. Extended X-ray absorption fine
`structure (EXAFS) studies [14] confirm the coordina-
`tion of iron by 6 oxygen atoms, at an Fe–O distance
`of 1.95 A˚ , and the location of a somewhat disordered
`shell of iron ions at a distance of about 3.05 A˚ . The
`iron oxyhydroxide crystallite dimensions estimated
`from the broadening of X-ray diffraction peaks are
`generally about 1–5 nm in diameter [9,14,17]. Electron
`microscopy has shown that the dimensions of the core
`(which may be larger than the size of the crystalline
`portion) can range from 3 nm diameter spheres [9]
`to ellipsoidal particles with dimensions of up to
`
`5 · 34 nm [21]. The mineral core is stabilized by inter-
`actions with the polysaccharide, presumably through
`the hydroxyl and/or carboxyl groups of the carbohy-
`drate units, the latter having been generated by partial
`oxidation of the polysaccharide. The overall particle
`size is highly dependent on the nature of the polysac-
`charide and the mode of complex formation. An older
`commercial
`iron–dextran complex had a reported
`overall diameter of about 12–13 nm [9], whereas iron
`complexes with j-carrageenan are reported to form
`25–50 nm aggregates of 10–20 nm particles [17,18].
`The apparent complex molecular mass, based on elu-
`tion position in gel permeation chromatography
`(GPC) chromatograms, may range from 72 kDa
`(ImferonÒ)
`(InFedÒ)
`to 90 kDa
`to 265 kDa
`(DexferrumÒ).
`Information about the structure of iron complexes
`with low molecular mass carbohydrates has been pre-
`dominantly derived from studies of soluble mono- or
`di-nuclear iron complexes. Weakly stable mononuclear
`complexes are formed by carbohydrates with struc-
`tures that allow at least
`three hydroxyl groups to
`interact simultaneously with a single iron ion. For
`example, sucrose or its component sugar fructose
`can bind iron weakly at neutral pH [4–6]. There is
`much less information about complexes containing
`polynuclear iron species complexed with monosaccha-
`rides or disaccharides. Studies of ferric fructose sys-
`tems suggested that the fructose may serve primarily
`to coat iron-containing particles at neutral pH. Inter-
`estingly, when the pH is raised, the formation of a tet-
`ra-deprotonated sugar moiety was proposed to break
`down the polynuclear iron to the mononuclear ferric
`fructose complexes. It was also noted that gluconate,
`with its carboxyl group, was much more effective in
`breaking down the particles [5]. For iron sucrose, ear-
`lier studies suggest the presence of polynuclear iron
`complexed with sucrose,
`in which the mineral core
`has been proposed to be a 2-line ferrihydrite [22]
`rather than an iron oxyhydroxide.
`Some of the questions that may arise in the analy-
`sis of polynuclear iron–carbohydrate complexes in-
`clude:
`(1)
`the composition and polymorphic form
`and degree of crystallinity of the iron mineral compo-
`nent, (2) the size and shape of the nanoparticles, (3)
`the stability of the nanoparticles, (4) the location of
`the saccharide component within the particle, (5) the
`molar ratio of iron to saccharide, and (6) the mode
`of binding between the iron and saccharide compo-
`nents,
`including the extent of saccharide deprotona-
`tion in the complex. In the present work, we address
`some of these questions by examining two different
`iron–carbohydrate complexes with respect to the min-
`eral environment of the iron, the nanoparticle size,
`and the particle stability upon dilution or separation
`from unbound carbohydrate by dialysis.
`
`Pharmacosmos A/S v. Luitpold Ex. Pharmaceuticals, Inc., IPR2015-01490
`
`Luitpold Pharmaceuticals, Inc., Ex. 2048, P. 2
`
`

`
`D.S. Kudasheva et al. / Journal of Inorganic Biochemistry 98 (2004) 1757–1769
`
`1759
`
`2. Materials and methods
`
`The parenteral iron formulations used in this study
`were an iron–sucrose (VenoferÒ, Iron Sucrose Injection,
`USP, lots 1644 and 2323A), from American Regent,
`Inc., Shirley, NY, and an iron–gluconate in sucrose
`(FerrlecitÒ, Sodium Ferric Gluconate Complex in Su-
`crose Injection, lots 1F541 and 1C239), from Watson
`Pharmaceuticals, Inc., Corona, CA. VenoferÒ was sup-
`plied as the iron–sucrose complex in 5 mL vials contain-
`ing 100 mg of elemental iron in 30% sucrose solution (20
`mg Fe/mL) at a pH of about 10.5–11.1. The apparent
`average molecular mass is stated in the package insert
`to be 34–60 kDa. FerrlecitÒ was obtained as the iron–
`gluconate complex in 5 mL ampules containing 62.5
`mg of elemental iron as a sodium salt of a ferric gluco-
`nate complex in 20% sucrose (12.5 mg Fe/mL) at a pH
`of about 7.7–9.7. The apparent average molecular mass
`is stated in the package insert to be 350 ± 23 kDa. Some
`samples, where noted in the text, were subjected to dial-
`ysis to remove low molecular mass components. Dialysis
`of samples for absorption spectroscopy was performed
`against deionized water, at a pH of about 5.4, in Slide-
`A-Lyzer Dialysis Cassettes (Pierce, Rockford, IL) made
`of low-binding regenerated cellulose membrane, molecu-
`lar weight cutoff (MWCO) 10,000. Cassette capacity was
`0.5–3 mL. Larger volumes of dialyzed samples were
`needed for AFM and TEM studies; in that case dialysis
`against deionized water at neutral pH, or with pH ad-
`justed with KOH to match the sample pH, was per-
`formed in tubes made of Spectra/Por Molecularporous
`Regenerated Cellulose Membrane (Spectrum Laborato-
`ries
`Inc., Rancho Domingues, CA) with MWCO
`6–8,000. Dialysis time for both trials was 20–48 h. Dia-
`lyzed samples were sent to Huffman Laboratories, Inc.
`(Golden, CO) for elemental analysis to determine the
`amount of iron, carbon and sodium.
`Absorption spectra were recorded on a Perkin–Elmer
`Lambda 800 UV/VIS absorption spectrophotometer.
`The spectra were analyzed with UV WinLab software
`v3.00.03. The spectra were recorded in the range of
`800–200 nm for original non-diluted samples and for di-
`luted or dialyzed samples. This instrument has the abil-
`ity to analyze samples with relatively high absorbance.
`Its photometric linearity at an absorbance (A) of 3
`is ± 0.006 A units and at A of 2 is ± 0.002 A units.
`The photometric accuracy at A of 2 is ± 0.003 A units
`and stray light
`in the range of 220–370 nm is
`<0.00008% T. The photometric range is ± 7 A. For
`the undiluted samples, cylindrical quartz cells of path
`length 0.001 or 0.01 cm were employed; for all other
`samples a cylindrical quartz cell of path length 0.01
`cm was used. The reference was water. All spectra were
`recorded as absorbance vs. wavelength of light in nm.
`Molar extinction coefficients were calculated from the
`absorbances at wavelengths of 300 and 470 nm, based
`
`on the molar concentration of iron in each sample.
`The molar iron concentrations were calculated from
`the stated weight concentration of iron in each sample.
`X-ray powder diffraction (XRD) patterns were re-
`1 in the range
`corded at a scanning rate of 0.008° min
`2h from 0° to 70° using a Philips X-ray diffractometer
`with Cu Ka radiation (k = 1.5418 A˚ ). Original and dia-
`lyzed samples were prepared by freeze-drying prior to
`analysis. The average iron mineral crystallite size, D,
`was calculated from the observed XRD profiles using
`the peak full width at half maximum height (b) of the se-
`lected diffraction peak, using the Scherrer formula [23]:
`D ¼ Kk=b cos h;
`where K is the Scherrer constant, with a value of 0.94.
`For recorded crystalline patterns, diffraction angles
`and interplanar spacings obtained were compared with
`the values compiled in the JCPDS-ICDD (Joint Com-
`mittee on Powder Diffraction Standard-International
`Center for Diffraction Data-copyright PSI Interna-
`tional) reference cards.
`Electron microscopic studies of the iron-core crystal
`size and morphology for the iron–carbohydrate samples
`were carried out by using a Phillips CM-12 Transmis-
`sion Electron Microscope (100 keV). All preparations
`were deposited one night before imaging onto a carbon
`stabilized Formvar-coated copper grid (400 mesh), pur-
`chased from Ted Pella, Inc., Redding, CA. Approximate
`grid hole size is 42 lm, and the thickness of Formvar
`films stabilized with carbon is 5–10 nm. All attempts
`to image the original undiluted liquids were not success-
`ful, because the resulting film on a grid was too thick for
`electrons to penetrate. Consequently, the iron–carbohy-
`drate complexes were dialyzed and the grid was dipped
`into the sample and then the excess liquid was removed
`with a very gentle flow of compressed nitrogen.
`Atomic force microscopy (AFM) was used to deter-
`mine size and shape of iron–carbohydrate complexes.
`The instrument used was a MultiMode Scanning Probe
`Microscope with Nanoscope IIIa controller and a type
`EV scanner (Veeco Instruments, Inc). The scanner was
`calibrated in the x–y plane using the 1 lm etched grid,
`and in the z-direction using the 200 nm height calibra-
`tion standard. The samples for imaging were prepared
`in two ways. In the first procedure, samples were diluted
`2 to 80-fold with deionized water prior to analysis. In
`the second procedure, samples were dialyzed against
`deionized water prior to analysis. In both cases, 4 lL
`of sample was put on a freshly cleaved mica surface,
`then after 20–30 s the surface was rinsed with 100 lL
`of deionized water to get rid of unbound particles, and
`the surface was dried by a gentle flow of nitrogen. The
`AFM imaging technique used was Tapping ModeTM.
`TESP etched silicon cantilever probes of 125 lm nomi-
`nal length were used, at a drive frequency of approxi-
`mately 240–280 kHz. We generally use an RMS (root
`
`Pharmacosmos A/S v. Luitpold Ex. Pharmaceuticals, Inc., IPR2015-01490
`
`Luitpold Pharmaceuticals, Inc., Ex. 2048, P. 3
`
`

`
`1760
`
`D.S. Kudasheva et al. / Journal of Inorganic Biochemistry 98 (2004) 1757–1769
`
`mean square) voltage of approximately 2.5–3 V, and ad-
`just the setpoint for optimum image, which is generally
`about 1 V below the RMS voltage. Both height and
`amplitude information were recorded at a scan rate of
`2–3 Hz, and stored in a 256 · 256 pixel format. Images
`were processed using the Nanoscope version 4.43r8 soft-
`ware. For optimum clarity in visual presentation of the
`nanoparticle images in the figures, flattening of first or-
`der was employed unless stated otherwise. For images to
`be used in measuring heights, the only image processing
`was zero order flattening. These images were analyzed to
`determine height and diameter of the particles observed.
`Although the height measurement is considered accu-
`rate, the apparent particle diameter is broadened by
`the finite dimensions of the probe tip. A correction for
`the tip broadening effect has been made following the
`procedure of Margeat et al. [24]. The apparent diameter
`measured at half height, S, is related to the real diame-
`ter, D, and the tip radius, R, according to the following
`
`equation [24]:
`S ¼ 2 RD þ D2
`4
`
`1=2
`
`:
`
`
`
`on the same day and with a single tip, and should be rep-
`resentative of the effect of dilution on the particle diam-
`eter. The relative sizes of particles in different iron
`formulations were confirmed by analysis of all samples
`on the same day, using a single tip, and running a cali-
`bration image of DNA on the same occasion.
`
`3. Results and discussion
`
`3.1. Absorption spectroscopy
`
`For Fe(III) in a high spin state octahedrally coordi-
`nated to oxygen, several characteristic absorption bands
`are expected. At 250–390 nm, a band of high intensity
`arises from oxo-metal charge transfer transitions, as
`has been reported for Fe(III)–glucose at 320 nm and
`for Fe(III)–sucrose at 370 nm [19]. Low intensity bands
`at 470–500 nm and 560–580 nm arise from d to d tran-
`sitions of octahedral d5 complexes of Fe(III). For low
`spin Fe(III), the latter bands would be of high intensity.
`The presence of Fe(II) would result in a band at 420–450
`nm for the d–d transition of octahedral d6 Fe(II), and a
`high intensity band at 800–900 nm for an aquated-Fe(II)
`complex [25].
`The spectral shapes of the studied parenteral iron for-
`mulations show a weak absorption band at about 470–
`500 nm and a strong band about 300 nm. Fig. 1 shows
`the absorption spectra for the two samples after dilution
`to the same nominal concentration of 5 mg Fe/mL, to
`facilitate spectral shape comparison. The 470–500 nm
`band is attributed to the d–d transition of octahedral
`high spin d5 Fe(III). The 300 nm band shows that the
`iron is chelated to oxygen, and is due to the oxo-metal
`
`The tip radius was estimated by imaging DNA, which
`has a real chain diameter of 2 nm. For one AFM tip
`used, the measured apparent diameter of DNA at half
`height was 11.8 nm, resulting in an estimated tip radius
`of 16.8 nm. A second tip had an estimated radius of 14.8
`nm. The appropriate tip radius was used to estimate cor-
`rected diameters in our images, although we may expect
`some minor variation for a given tip as a function of use.
`Thus the estimated particle diameters are subject to
`uncertainty on this basis. The relative diameters for a gi-
`ven series of samples differing in dilution were measured
`
`4.0
`
`3.5
`
`3.0
`
`2.5
`
`2.0
`
`1.5
`
`1.0
`
`0.5
`
`Absorbance
`
`0.0
`250
`
`300
`
`350
`
`400
`
`550
`500
`450
`Wavelength (nm)
`
`600
`
`650
`
`700
`
`750
`
`800
`
`Fig. 1. Absorption spectra of iron–sucrose (dashed line), and iron–gluconate (solid line) diluted to nominal iron concentrations of 5 mg/mL and
`analyzed in an optical cell of 0.01 cm path length.
`
`Pharmacosmos A/S v. Luitpold Ex. Pharmaceuticals, Inc., IPR2015-01490
`
`Luitpold Pharmaceuticals, Inc., Ex. 2048, P. 4
`
`

`
`D.S. Kudasheva et al. / Journal of Inorganic Biochemistry 98 (2004) 1757–1769
`
`1761
`
`Table 1
`Characterization of iron–carbohydrate complexes by absorption spectroscopy
`
`Sample
`
`Iron–sucrose
`
`Iron–gluconate
`
`Fe Conc. (mg/mL)
`
`Cell path length (cm)
`
`e300 (M
`
`1 cm
`
`1)
`
`e470 (M
`
`1 cm
`
`1)
`
`e300/e470
`
`20
`20
`5
`
`12.5
`12.5
`5
`
`0.001
`0.01
`0.01
`
`0.001
`0.01
`0.01
`
`2990
`
`2720
`
`2600
`
`2670
`
`380
`390
`
`290
`250
`
`7.9
`
`7.0
`
`9.0
`
`10.6
`
`charge transfer absorption band. There is no evidence
`for the existence of low spin state Fe(III) or for aquated
`Fe(II). The shapes of the spectra are essentially the same
`for both iron–carbohydrate samples; both formulations
`showed characteristic absorption spectra expected for
`Fe(III) complexes in an octahedral d5 high spin state.
`The position of the d–d transition band around 500
`nm has been shown to depend on the distance between
`two neighboring Fe atoms [26]. Thus, in the case of
`face-sharing octahedra of an iron oxide hematite, this
`band is above 500 nm. For the edge- and corner-sharing
`octahedra and hence larger Fe–Fe distances in iron oxy-
`hydroxides this transition is weaker and lies at shorter
`wavelengths of 470–499 nm [27]. Based on the fact that
`both iron formulations have this band at approximately
`the same position, we can conclude that the nature of
`the iron core is the same and both spectra are compati-
`ble with the expected ferric oxyhydroxide mineral form.
`The molar extinction coefficients of the absorption
`bands for each sample were closely similar (Table 1).
`The peak at 300 nm was best quantitated in undiluted
`samples analyzed in an optical cell with a path length
`of 0.001 cm. Calculated extinction coefficients at 300
`1 cm
`1 for the
`nm were approximately 2600–3000 M
`two samples. This compares with reported values of
`6660 for a different iron–sucrose preparation [19] and
`1700 for iron–carrageenan [17]. The shoulder at 470–
`
`500 nm was best quantitated in undiluted sample ana-
`lyzed in an optical cell with a 0.01 cm path length.
`The extinction coefficient at 470 nm was approximately
`1 cm
`1 for the two samples.
`300–400 M
`The stability of the iron core structure in each of the
`iron formulations was assessed by dilution and/or dialy-
`sis. Dilution to a concentration of 5 mg Fe/mL, or dial-
`ysis against water resulting in a similar dilution, had no
`significant effect on the spectral shape or molar extinc-
`tion coefficients of the iron absorption bands (Table 1).
`
`3.2. Transmission electron microscopy (TEM)
`
`TEM micrographs of dialyzed iron–sucrose and dia-
`lyzed iron–gluconate samples at equivalent magnifica-
`tions are compared in Fig. 2. The beadlike iron
`mineral cores appear to be fairly uniform in size for a gi-
`ven sample. This uniformity reflects the manufacturing
`processes, and results in particles with a narrow range
`of sizes. The comparison of the images of two samples
`shows that the iron–gluconate particles have cores that
`are generally slightly smaller (Fig. 2(a)) than those of
`iron–sucrose (Fig. 2(b)).
`Measurement of the diameters of the electron dense
`particles is used to estimate the diameter of the iron min-
`eral core in a given sample. For iron–sucrose, the results
`indicate a mineral core size distribution ranging from
`
`Fig. 2. Transmission electron microscopy image of (a) dialyzed iron–gluconate complex and (b) dialyzed iron–sucrose complex. The bar is 20 nm.
`
`Pharmacosmos A/S v. Luitpold Ex. Pharmaceuticals, Inc., IPR2015-01490
`
`Luitpold Pharmaceuticals, Inc., Ex. 2048, P. 5
`
`

`
`1762
`
`D.S. Kudasheva et al. / Journal of Inorganic Biochemistry 98 (2004) 1757–1769
`
`approximately 1.0–6.5 nm. The average size of the core is
`3 ± 2 nm (n = 20). For iron–gluconate, the mineral core
`diameter ranges from approximately 0.9–3.5 nm, and
`the average diameter of the core is 2 ± 1 nm (n = 20).
`
`3.3. X-ray powder diffraction
`
`The chemical structure of the iron core was evaluated
`by means of X-ray diffraction (XRD) analysis. XRD is
`considered to be the most reliable way to identify a par-
`ticular iron polymorph because it is based on the long
`range order of the atoms [3]. The diffractograms ob-
`
`tained for lyophilized samples of original, nondialyzed
`preparations were compared with reference data for
`different iron oxyhydroxide mineral phases, similar in
`composition and structure (Table 2). Both iron–
`carbohydrate samples give a crystalline pattern of a sin-
`gle phase whose peak positions and intensities match
`with reported JSPDS-ICDD reference data for iron
`(III) oxyhydroxide akaganeite b-FeOOH. The X-ray re-
`sults are not consistent with the ferrihydrite structure
`found in ferritin [28]. In addition, two peaks in the dif-
`fractograms of both samples in the small angle region
`(11.5° and 20°) belong to sucrose. From the analysis
`
`Table 2
`Interplanar spacings in Angstroms and relative strength of reflections (from JCPDS-ICDD cards for sucrose and iron minerals and from recorded
`difractograms for original nondialyzed formulations)
`
`Iron–sucrose (angle)
`7.59 (11.5°)
`
`Iron–gluconate (angle)
`7.54 (11.7°)
`
`6.91 (13.1°)
`
`4.68 (17.1°)
`
`3.61 (22.3°)
`
`3.61 (22.3°)
`
`2.64 (33.5°)
`
`2.56 (35.1°)
`
`2.34 (38.3°)
`2.28 (39.6°)
`
`1.76 (51.8°)
`
`1.64 (55.5°)
`1.52 (61.1°)
`
`1.49 (62.3°)
`
`2.69 (33.3°)
`
`2.555 (35.13°)
`
`2.35 (38.2°)
`
`1.75 (52.01°)
`
`1.66 (55.2°)
`
`1.50 (61.8°)
`1.49 (62.3°)
`
`Sucrose
`
`7.58s
`
`6.94m
`6.73m
`
`4.71s
`4.52s
`4.25m
`
`3.59s
`3.53m
`
`2.88w
`
`2.35m
`2.339w
`
`2.028w
`1.971w
`
`Hematite
`
`Goethite
`
`Ferrihydrite
`
`Akaganeite
`
`7.467s
`
`5.276m
`
`3.728w
`
`3.333s
`
`2.634m
`
`2.550s
`
`2.356w
`2.295m
`
`1.954m
`
`1.755w
`
`1.643m
`1.515w
`1.503w
`1.489w
`1.485w
`1.445m
`
`4.98w
`
`4.18s
`
`2.69m
`
`2.58w
`
`2.45s
`
`2.25w
`2.19m
`
`2.50s
`
`2.21s
`
`1.96s
`
`1.72m
`
`1.72m
`
`1.51m
`
`1.48s
`
`1.42w
`1.393w
`1.36w
`1.345w
`1.317w
`1.29w
`
`3.68m
`
`2.70s
`
`2.52s
`
`2.21m
`
`1.84m
`
`1.69s
`
`1.49m
`
`1.45m
`
`1.055w
`1.033w
`0.960w
`0.9581w
`0.8436w
`
`s, strong; m, medium; and w, weak.
`
`Pharmacosmos A/S v. Luitpold Ex. Pharmaceuticals, Inc., IPR2015-01490
`
`Luitpold Pharmaceuticals, Inc., Ex. 2048, P. 6
`
`

`
`D.S. Kudasheva et al. / Journal of Inorganic Biochemistry 98 (2004) 1757–1769
`
`1763
`
`220
`
`200
`
`180
`
`160
`
`140
`
`120
`
`100
`
`80
`
`60
`
`40
`
`20
`
`Intensity
`
`5
`
`15
`
`25
`
`35
`Degrees 2-Theta
`
`45
`
`55
`
`65
`
`Fig. 3. XRD patterns of iron–sucrose (top) and iron–gluconate (bottom).
`
`of the diffractograms in Fig. 3, it can be also concluded
`that all peaks broadened evenly, without the presence of
`a selective peak broadening, i.e., there was no preferen-
`tial deformation plane due to the complexation with su-
`crose or gluconate.
`It should be mentioned that XRD patterns of both
`formulations appear to have a very weak diffraction
`peak
`from the
`(0 0 2)
`plane
`(diffraction
`angle
`
`2h = 61.15°), which would indicate the presence of Cl
`
`anions. The presence of Cl
`is in accord with the tun-
`nel-containing akaganeite structure.
`On the XRD pattern of iron–sucrose sample after
`pH-matched dialysis, the sucrose peaks at 11.5° and
`22° are almost eliminated (Fig. 4, top spectrum). How-
`
`lifting of the area of their presence
`ever, an overall
`around 20° confirms that there is still some sucrose in
`the sample and that the sucrose is not uniformly or-
`dered. Dialysis did not destroy the iron mineral core,
`since the main reflection attributed to the akaganeite
`polymorph remains strong after dialysis.
`For iron–gluconate, the effect of removing unbound
`sucrose was more noticeable (Fig. 4, bottom spectrum).
`In addition to the apparent disappearance of the sucrose
`peaks at 11.5° and 22°, the pattern of the dialyzed sam-
`ple has a newly visible akaganeite peak at 26.8°. That
`difference in the peak appearance between the two dia-
`lyzed samples occurs despite the fact that the size of
`the iron–gluconate core is slightly smaller than that of
`
`200
`
`180
`
`160
`
`140
`
`120
`
`100
`
`Intensity
`
`80
`
`60
`
`40
`
`20
`
`5
`
`15
`
`25
`
`35
`Degrees 2-Theta
`
`45
`
`55
`
`65
`
`Fig. 4. XRD patterns of dialyzed iron–sucrose (top) and dialyzed iron–gluconate (bottom).
`
`Pharmacosmos A/S v. Luitpold Ex. Pharmaceuticals, Inc., IPR2015-01490
`
`Luitpold Pharmaceuticals, Inc., Ex. 2048, P. 7
`
`

`
`1764
`
`D.S. Kudasheva et al. / Journal of Inorganic Biochemistry 98 (2004) 1757–1769
`
`iron–sucrose, as demonstrated by TEM. On the basis of
`overall core size, the peaks in the XRD pattern of iron–
`gluconate should be less pronounced. Thus, the differ-
`ence in the diffractogram appearance may indicate more
`uniform crystallites in the iron–gluconate than in the
`iron–sucrose, after dialysis to remove unbound and
`loosely bound carbohydrate.
`Our FTIR studies (unpublished data) of different iron
`oxyhydroxide polymorphs also confirmed that the iron
`core of both formulations has the chemical structure
`of iron (III) oxyhydroxide akaganeite b-FeOOH.
`The average crystallite size, D, was estimated from
`the diffraction peak from the (2 1 1) plane (diffraction an-
`gle 2h = 35.1°), and gave a similar result for both formu-
`lations.
`In the case of nondialyzed samples,
`the
`crystallite size for iron–sucrose was estimated to be 4.6
`nm, while for iron gluconate it was 4.1 nm. For the
`iron–sucrose sample after pH-adjusted dialysis (Fig. 4),
`D calculated for the same peak gave a smaller size of
`3.4 nm. The average crystallite size for the dialyzed
`iron–gluconate was 3.5 nm. These values are slightly
`higher than the core diameters estimated for dialyzed
`samples by transmission electron microscopy. Consider-
`ing the difficulty in accurately measuring the diffraction
`peak width at half height, the TEM data are considered
`more reliable.
`
`3.4. Atomic force microscopy
`
`Imaging the iron–carbohydrate complexes was per-
`formed for samples applied to a mica surface, followed
`by brief rinsing to remove unbound material, including
`excess soluble low molecular mass carbohydrates. The
`images of the two iron formulations showed large num-
`bers of beadlike nanoparticles spread over the scanning
`surface.
`The arrangement of particles on the mica surface is
`different from the arrangement on the surface of the
`TEM grid, which is quite expected because the surfaces
`are different. AFM mica is hydrophilic, whereas the
`TEM grid is covered with carbon, which makes it highly
`hydrophobic. Thus the particles are more uniformly dis-
`tributed on the mica surface then on the carbon coated
`TEM grid.
`Fig. 5 shows 3 lm · 3 lm scans of each iron formu-
`lation sample at an iron concentration of 5 mg/mL.
`The surfaces showed residual excess sucrose or sucrose
`and gluconate surrounding the nanoparticles. The bead-
`like particles, which include both the mineral core and
`the bound carbohydrate, appear to be similar in size
`for iron–sucrose and iron–gluconate.
`Images obtained at higher resolution are preferred for
`more accurate measurements of the nanoparticle sizes.
`Fig. 6 shows 1 lm · 1 lm scans of nanoparticles from
`each preparation type, analyzed at comparable iron con-
`centrations after dilution with water. The distributions
`
`Fig. 5. Atomic force microscope height images of: (a) iron–sucrose
`complex, (b) iron–gluconate complex in sucrose, after dilution to an
`iron concentration of 5 mg/mL. Image size 3 lm · 3 lm, with a color
`scale covering a height range of 15–20 nm (darker features being lower
`in height than lighter ones).
`
`of particle heights measured from the images in Fig. 6
`are given in Table 3 for iron–sucrose and Table 4 for
`iron–gluconate. An equal number of particles (75) were
`analyzed at each concentration for each preparation
`type.
`At a concentration of about 5 mg Fe/mL, the iron–
`sucrose complexes (Fig. 6(a) and Table 3) appear as
`beads with an average height of about 3.5 nm, and a cor-
`rected width of approximately 22 nm. Analyzed at the
`same 5 mg Fe/mL concentration, the iron–gluconate in
`sucrose complex (Fig. 6(e) and Table 4) had an average
`height of approximately 3.1 nm, and an average appar-
`ent particle width of approximately 12 nm.
`For an incompressible spherical bead, the height and
`width measurements should both be equal to the bead
`diameter. Two factors contribute to the mismatch be-
`tween vertically measured height and horizontally meas-
`ured diameter. The first is a flattening effect caused by
`the tapping impact of the probe. The carbohydrate com-
`ponent of the beads is subject to compression and dis-
`
`Pharmacosmos A/S v. Luitpold Ex. Pharmaceuticals, Inc., IPR2015-01490
`
`Luitpold Pharmaceuticals, Inc., Ex. 2048, P. 8
`
`

`
`D.S. Kudasheva et al. / Journal of Inorganic Biochemistry 98 (2004) 1757–1769
`
`1765
`
`observed upon dilution to 2 mg Fe/mL, was primarily
`caused by the fact that the particles were deeply embed-
`ded in a sucrose residue covering the surface for that
`particular sample (Fig. 6(b)). Thus, the measured height
`in that case is really only the height of the upper part of
`the particles above the sucrose. That phenomenon can
`also explain the observation that upon the next dilution
`to 0.8 mg/mL the average measured height increased
`again to 3 nm (Fig. 6(c)), in contrast with the logical
`expectation that upon dilution the particle size would
`decrease. Upon the final dilution to 0.66 mg/mL the
`average measured height stayed nearly constant at 3
`nm. We conclude that the height of the particles is rela-
`tively stable with dilution. The fact tha

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