throbber
European Journal of Pharmaceutics and Biopharmaceutics 78 (2011) 480–491
`
`Contents lists available at ScienceDirect
`
`European Journal of Pharmaceutics and Biopharmaceutics
`
`j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / e j p b
`
`Research paper
`A comparative study of the physicochemical properties of iron isomaltoside
`1000 (MonoferÒ), a new intravenous iron preparation and its clinical implications
`Markus R. Jahn a, Hans B. Andreasen b, Sören Fütterer a, Thomas Nawroth a, Volker Schünemann c, Ute Kolb d,
`Wolfgang Hofmeister e, Manuel Muñoz f, Klaus Bock g, Morten Meldal g, Peter Langguth a,⇑
`
`a Institute of Pharmacy and Biochemistry, Johannes-Gutenberg University, Mainz, Germany
`b Pharmacosmos A/S, Holbaek, Denmark
`c Department of Physics, Technical University Kaiserslautern, Kaiserslautern, Germany
`d Institute of Physical Chemistry, Johannes-Gutenberg University, Mainz, Germany
`e Institute of Geosciences, Johannes-Gutenberg University, Mainz, Germany
`f Transfusion Medicine, School of Medicine, University of Málaga, Málaga, Spain
`g Carlsberg Laboratory, Gamle Carlsberg Vej 10, Valby, Denmark
`
`a r t i c l e
`
`i n f o
`
`a b s t r a c t
`
`Article history:
`Received 14 December 2010
`Accepted in revised form 15 March 2011
`Available online 23 March 2011
`
`Keywords:
`Inorganic nanoparticles
`Colloids
`Stability
`Free iron
`Iron supplementation
`Release rate
`
`The treatment of iron deficiency anemia with polynuclear iron formulations is an established therapy in
`patients with chronic kidney disease but also in other disease areas like gastroenterology, cardiology,
`oncology, pre/post operatively and obstetrics’ and gynecology. Parenteral iron formulations represent
`colloidal systems in the lower nanometer size range which have traditionally been shown to consist of
`an iron core surrounded by a carbohydrate shell. In this publication, we for the first time describe the
`novel matrix structure of iron isomaltoside 1000 which differs from the traditional picture of an iron core
`surrounded by a carbohydrate. Despite some structural similarities between the different iron formula-
`tions, the products differ significantly in their physicochemical properties such as particle size, zeta
`potential, free and labile iron content, and release of iron in serum. This study compares the physiochem-
`ical properties of iron isomaltoside 1000 (MonoferÒ) with the currently available intravenous iron prep-
`arations and relates them to their biopharmaceutical properties and their approved clinical applications.
`The investigated products encompass low molecular weight iron dextran (CosmoFerÒ), sodium ferric glu-
`conate (FerrlecitÒ), iron sucrose (VenoferÒ), iron carboxymaltose (FerinjectÒ/InjectaferÒ), and ferumoxy-
`tol (FerahemeÒ) which are compared to iron isomaltoside 1000 (MonoferÒ). It is shown that significant
`and clinically relevant differences exist between sodium ferric gluconate and iron sucrose as labile iron
`formulations and iron dextran, iron carboxymaltose, ferumoxytol, and iron isomaltoside 1000 as stable
`polynuclear formulations. The differences exist in terms of their immunogenic potential, safety, and con-
`venience of use, the latter being expressed by the opportunity for high single-dose administration and
`short infusion times. Monofer is a new parenteral iron product with a very low immunogenic potential
`and a very low content of labile and free iron. This enables Monofer, as the only IV iron formulation,
`to be administered as a rapid high dose infusion in doses exceeding 1000 mg without the application
`of a test dose. This offers considerable dose flexibility, including the possibility of providing full iron
`repletion in a single infusion (one-dose iron repletion).
`
`Ó 2011 Elsevier B.V. All rights reserved.
`
`1. Introduction
`
`Parenteral iron therapy is today widely used for the treatment
`of iron deficiency anemia. Patients with chronic kidney disease
`(CKD) also frequently need treatment with parenteral iron prepa-
`rations in addition to erythropoietin stimulating agents [1]. For
`
`⇑ Corresponding author. Institute of Pharmacy and Biochemistry, Pharmaceutical
`
`Technology and Biopharmaceutics, Johannes-Gutenberg University Mainz, D-55099
`Mainz, Germany.Tel.: +4961313925746; fax: +49 6131 3925021.
`E-mail address: langguth@uni-mainz.de (P. Langguth).
`
`0939-6411/$ - see front matter Ó 2011 Elsevier B.V. All rights reserved.
`doi:10.1016/j.ejpb.2011.03.016
`
`renal failure patients on dialysis, the average iron requirements
`due to blood loss are equivalent to 1–3 g of elemental iron per year
`[2]. This can easily be accomplished by frequent low dose IV iron
`administrations, during the regular dialysis sessions.
`From initial, generalized use in nephrology parenteral iron ther-
`apy has spread in recent years to other disease areas; gastroenter-
`ology [3], cardiology [4,5], oncology [6], pre/post operatively [7],
`obstetrics’, and gynecology [8]. However, care providers in these
`segments have less frequent patient contact, resulting in an in-
`creased demand for convenient administration of large IV iron
`doses in one clinical session.
`
`Pharmacosmos, Exh. 1029, p. 1
`
`

`
`M.R. Jahn et al. / European Journal of Pharmaceutics and Biopharmaceutics 78 (2011) 480–491
`
`481
`
`Historically, the first parenteral iron preparations were toxic,
`being administered as an iron oxyhydroxide complex. This prob-
`lem was circumvented with the introduction of compounds con-
`taining iron in a core surrounded by a carbohydrate shell [9]. The
`currently marketed parenteral iron preparations are considered
`equally efficacious but vary in molecular size, pharmacokinetics,
`and adverse reaction profiles. The intravenous iron agents cur-
`rently available include high molecular weight iron dextran
`low molecular weight iron dextran (CosmoferÒ,
`(DexferrumÒ),
`InfedÒ), sodium ferric gluconate (FerrlecitÒ), iron sucrose (VenoferÒ),
`iron carboxymaltose (FerinjectÒ/InjectaferÒ), and ferumoxytol (Fera-
`hemeÒ). High molecular weight iron dextran has been linked to an
`increased risk of anaphylaxis and anaphylactoid reactions, and it is
`not available in Europe [10–13]. Although this problem is very
`much reduced with low molecular weight iron dextran [10–13],
`there is still a test dose requirement and the infusion of larger
`doses is hampered by a 4–6 h infusion time. Sodium ferric gluco-
`nate and iron sucrose can only be used in moderate iron doses
`due to the relative weakness of the iron complex [14]. Two new
`parenteral iron compounds, iron carboxymaltose, and ferumoxytol
`were recently introduced in the EU and the US markets, respec-
`tively. The FDA failed to approve iron carboxymaltose for distribu-
`tion in the USA due to unexplained hypophosphatemia, an
`increased number of adverse cardiac events and an imbalance in
`death rates in the treatment arm compared to the control arm in
`different RCTs [15].
`Although more stable than sodium ferric gluconate and iron su-
`crose, the administration of iron carboxymaltose and ferumoxytol
`is still limited to a maximum total dose of 1000 mg and 510 mg,
`respectively.
`The newest IV iron agent Iron isomaltoside 1000 (MonoferÒ)
`(e.g., iron oligo isomaltoside (1000) as generic name) is developed
`and manufactured by Pharmacosmos in Denmark and was intro-
`duced in Europe in 2010. The carbohydrate isomaltoside 1000 is
`a pure linear chemical structure of repeating a1-6 linked glucose
`units, with an average size of 5.2 glucose units and an average
`molecular weight of 1000 Da, respectively. It is a nonbranched,
`nonanaphylactic carbohydrate [16,17], structurally different from
`branched polysaccharides used in iron dextran (Cosmofer).
`The production method and the short nonionic isomaltoside
`1000 allows for the construction of a special matrix-like structure
`with interchanging iron molecules and linear isomaltoside 1000
`oligomers. The resulting matrix contains about 10 iron molecules
`per one isomaltoside pentamer in a strongly bound structure that
`enables a controlled and slow release of bioavailable iron to iron-
`binding proteins with little risk of free iron toxicity [18,19]. This al-
`lows iron isomaltoside 1000 to be administered safely as a rapid
`high dose intravenous infusion or bolus injection offering consider-
`able dose flexibility, including the possibility of providing full iron
`repletion in a single infusion, the so-called one-dose iron repletion.
`This article introduces and compares physicochemical proper-
`ties of iron isomaltoside 1000 (MonoferÒ) with currently marketed
`iron formulations. In addition, this comparative study of polynu-
`clear iron formulations currently used in the treatment of anemic
`disorders includes perspectives on the relevance of these properties
`with respect to safety, efficacy, and convenience of administration.
`
`2. Materials and methods
`
`2.1. Materials
`
`mL in 2 mL ampoules; Teva, Mörfelden-Walldorf, Germany), iron
`isomaltoside 1000 (MonoferÒ, 100 mg Fe/mL in vials; Pharmacos-
`mos, Holbaek, Denmark), iron carboxymaltose (FerinjectÒ, 50 mg
`Fe/mL in 2 mL vials; Vifor, München, Germany), and ferumoxytol
`(FerahemeÒ, 30 mg Fe/mL, in 17 mL vials; AMAG Pharmaceuticals,
`Lexington, MA, USA) were obtained from a pharmacy or directly
`from the manufacturer. The FerrozineÒ reaction kit was purchased
`from Roche Diagnostics GmbH, Mannheim. All iron formulations
`were used immediately after opening the vial or kept at 4 °C under
`nitrogen. Solutions were made from double-distilled water.
`
`2.2. Gel permeation chromatography (GPC)
`
`The apparent average molecular weight was analyzed by gel
`permeation chromatography. Prior to sample analysis, the columns
`were calibrated using dextran standards. The dextran standards
`used for GPC calibration were the commercial available Pharma-
`cosmos standards and consisted of Dextran 25, 50, 80, 150, 270,
`and 410, respectively. The average molecular weights Mw and the
`peak average molecular weights MP were 23.000, 21.400; 48.600,
`43.500; 80.900, 66.700; 147.600, 123.600; 273.000, 196.300;
`409.800, 276.500 for Dextran 25, 50, 80, 150, 270, and 410, respec-
`tively. The standards have been evaluated against the Ph.EUR and
`USP dextran standards.
`The detector used in the GPC measurements is a VE 3580 RI
`detector (Viscotec). Data are collected and calculations are made
`using the Omnisec 4.1 software from Viscotec.
`The hydrodynamic diameter dh was calculated from the hydro-
`dynamic volume Vh = Mp |g|, where the intrinsic viscosity |g| is
`given by the Mark Houwink equation [20]
`jgj ¼ kMa
`v
`
`where Mav is the viscosity average molecular weight.
`
`2.3. Dynamic light scattering (DLS) and zeta potential
`
`The size distribution and zeta potential of the whole particle,
`which can include an iron hydroxide core plus a carbohydrate
`shell, was determined by DLS. The diluted samples (0.4 mg Fe/mL
`double-distilled and sterile filtered water) were measured using
`a Zetasizer Nano S (Malvern Instruments Ltd.; Worcestershire,
`UK) including a He–Ne Laser with a wavelength of k = 633 nm,
`which illuminated the samples and detects the scattering informa-
`tion at an angle of 173° (Noninvasive Back-scatter technology).
`Zeta potential measurements were performed at different pH val-
`ues by addition of 0.1 N HCl or NaOH, respectively. The data were
`analyzed with the firmware, Zetasizer Software DTSv612 yielding
`volume distribution data.
`
`2.4. Transmission electron microscopy (TEM)
`
`The dimension of the iron complex nanoparticle core was deter-
`mined with an EM420 transmission electron microscope (FEI/Phi-
`lips, Oregon, USA) at 120 kV. All preparations (1 mg Fe/mL,
`double-distilled water) were deposited onto a hydrophilized cup-
`
`ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiq
`per grid (300 mesh, Ø 3 mm) and were allowed to dry. The median
`
`of the geometrical diameter dg ¼
`
`ðd2s þ d2l Þ=2
`was determined
`(n = 50, ds = shortest dimension, dl = longest dimension).
`
`2.5. X-ray diffraction (XRD)
`
`Sodium ferric gluconate (FerrlecitÒ, 12.5 mg Fe/mL in 3.2 mL
`ampoules; Sanofi-Aventis, Frankfurt, Germany),
`iron sucrose
`(VenoferÒ, 20 mg Fe/mL in 5 mL ampoules; Vifor, München, Ger-
`many), low molecular weight iron dextran (CosmoFerÒ, 50 mg Fe/
`
`X-ray measurements of dried out solutions (30 °C) were per-
`formed with a XRD 3000 TT (Seifert, Ahrensburg, Germany) using
`Cu radiation (k = 1,54178 Å, 40 kV, 30 mA) in Bragg Brentano con-
`figuration (automatic divergence slit, angular rate 0,18°/min).
`
`Pharmacosmos, Exh. 1029, p. 2
`
`

`
`482
`
`M.R. Jahn et al. / European Journal of Pharmaceutics and Biopharmaceutics 78 (2011) 480–491
`
`The particles mean diameter d was determined from the Scher-
`rer equation: d ¼ k
`b cos h, where b is the full width at half maximum
`of the peak at 36° 2h or 63° 2h.
`
`2.6. Mössbauer spectroscopy
`
`Mössbauer spectra of iron isomaltoside 1000 were recorded
`using a conventional spectrometer in the constant-acceleration
`mode. Isomer shifts are given relative to a-Fe at room temperature.
`The spectra were measured in a closed cycle cryostat (Cryo
`Industries of America, USA) at 150 K, equipped with permanent
`magnets. The magnetically split spectra were analyzed by least-
`square fits using Lorentzian line.
`
`2.7. Dialysable iron in buffer
`
`The amount of free iron was estimated using the dialysis tech-
`nique following pH adjustment of each iron dispersion to 7.5. A dis-
`persion volume containing 150 mg of iron (7.5 mL for LMW iron
`dextran, iron isomaltoside 1000, iron carboxymaltose and ferum-
`oxytol, respectively; 15.0 mL for sodium ferric gluconate, and
`11.25 mL for iron sucrose) was added resulting in concentrations
`of 20.0 mg Fe/mL for all iron products except for sodium ferric gluco-
`nate (10.0 mg Fe/mL) and iron sucrose (13.3 mg Fe/mL). Dilutions
`were made with water and 0.9% sodium chloride solution, respec-
`tively. The volumes were added inside the dialysis tubing (12,000–
`14,000 MWCO, Medicell, London, United Kingdom) and dialyzed
`for 24 h at 20 °C against 100 mL of water or sodium chloride solution,
`respectively. The total volume including the dialysis tube was
`107.5 mL. Dialysis of each iron agent was performed in duplicate.
`Iron in the surrounding solution was quantified using ICP-MS (induc-
`tively coupled plasma mass spectrometry). The ICP-MS instrument
`was a Thermo iCap 6000 ICP-OES (Thermo Scientific, Danmark).
`Iron is measured at 238,201 nm. The measurement is made ax-
`ial. Two-point (left–right) baseline correction and external linear
`calibration curve are used.
`The experiments were carried out at room temperature (20–
`24 °C). In order to evaluate the effect of pH on the level of dialysable,
`free iron above experiments were conducted also for the high dose IV
`iron formulations low molecular weight iron dextran, iron isomalt-
`ose, iron carboxymaltose, and ferumoxytol without pH adjustment.
`
`2.8. Acid soluble FeOOH
`
`The acidic hydrolysis of the FeOOH in [FeOOH]mLn was followed
`by quantifying the decreasing FeOOH concentration with UV-
`spectroscopy. The spectrometer used was a Lambda 20 (Perkin
`Elmer). Readings at 287.3 nm were made from a scan using data
`interval 1.0 nm, scan speed 249 nm/min, a slit with of 2.0 nm
`and a smooth with of 2.0 nm.
`The absorbance of iron agents (10 mg Fe/l, 10 mm path length)
`in 0.9% NaCl/0.2375 M HCl was measured at 287.3 nm from
`t = 0 min to t = 48 h, unless otherwise specified. Initial absorbance
`after dilution of the iron preparation at t  0 min was set to 1
`according to 100% undissolved FeOOH and all other measurements
`were normalized for this. Ln(normalized data) was plotted against
`time and fitted with a second degree polynomial (R2 > 0.990). Half-
`life t0.5 was calculated from fpolynomial(t0.5) = ln(0, 5).
`
`2.9. FerrozineÒ-detectable labile iron
`
`The dissolution of iron in serum was determined by the Ferro-
`zineÒ-method [21–24]. FerrozineÒ does not only detect the free
`iron but also the weakly bound iron in the complex and the trans-
`ferrin bound iron in serum, this determination allows one to quan-
`tify the in vitro labile iron pool of the investigated intravenous iron
`
`formulations. By this method, iron is detected in the ferrous as well
`as the ferric state as the ferric iron is reduced by ascorbate to fer-
`rous iron. Briefly, human serum was incubated with the iron prep-
`aration corresponding to theoretical doses of 200 mg and 500 mg
`leading to a serum concentration of 66.7 lg/mL and
`iron,
`166.7 lg/mL for a person with a body mass of 70 kg, respectively.
`These serum concentrations are consequences of a blood volume
`of 0.07 L per kg and a serum fraction of 60% of the blood volume,
`yielding a total serum volume of approx. 3 L [25]. The experiment
`was performed at room temperature (22 °C) in 1.5 mL Eppendorf-
`tubes. Incubations were done for 10 and 45 min, respectively.
`Thereafter, a 100 lL sample was analyzed by addition of 700 lL re-
`agent 1 containing thiourea (115 mM) and citric acid (200 mM),
`followed by addition of 350 lL of reagent 2 containing sodium
`ascorbate (150 mM) and FerrozineÒ (6 mM). Absorption of the
`complex was measured at 562 nm over approximately 60 min
`using a PERKIN ELMER Lambda 20 (Perkin Elmer Inc., Waltham,
`MA, USA) UV–Vis spectrometer. The obtained absorbance versus
`time curve was fitted to a second degree polynomial for each incu-
`bation period and the intercept with the ordinate was calculated to
`receive the comparable theoretical amount of FerrozineÒ-detect-
`able iron. The regression coefficient for the polynomial function
`was always better than 0.995. The labile iron pool was calculated
`by linear regression analysis of the obtained intercepts from curves
`at 10 min incubation and 45 min incubation.
`
`2.10. Elucidation of molecular structure of iron isomaltoside 1000
`
`Proton and carbon NMR spectra were obtained on a Bruker
`800 MHz NMR instrument as ca. 5% solutions in D2O at 300 K. Sig-
`nals were referenced to external dioxane.
`The iron isomaltoside 1000 formulation (10.3 mg) was dis-
`solved in D2O (600 lL). The sample was transferred to a 5 mm
`NMR-tube and the 13C NMR spectrum was recorded at 20 °C on a
`Bruker Avance 800 instrument at 201.12 MHz for carbon
`(799.96 MHz for proton), integrated and compared with the spec-
`trum of the oligosaccharide alone (7.3 mg) in D2O (600 lL) [26].
`Both samples were measured again and the signals integrated after
`addition of 2.24 mg of methyl b-maltoside as internal reference.
`Molecular modeling: First the isomaltodisaccharide was con-
`structed and an MD calculation using the modeling program,
`MOE (Molecular Operating Environment, Version 2009.10, Chemi-
`cal Computing Group Inc., Montreal, Canada), at 450 K, stepsize
`0,1 fs clearly showed a significant preference for the gt conforma-
`tion of C-5–C-6 bonds independent of the starting point. The O-1–
`C-6 of the glycosidic bond had a weak preference for a trans-
`arrangement and the orientation of the C-1–O-1 bond satisfied
`the exoanomeric effect.
`The isomaltoside pentamer (composed of 5 a1-6 linked glucose
`molecules) with glucitol at the reducing end was built from disac-
`charides in their preferred conformation and energy minimized.
`The molecule was soaked in water (eight layers) and molecular
`dynamics was performed at the above conditions corresponding
`to a period of 2 ns. The additive effect of the oligosaccharide repeat
`to stabilize the preferred conformation when compared to the
`disaccharide was significant. The resulting structure was re-soaked
`and was subjected to energy minimization in water.
`
`3. Results
`
`3.1. Overall particle size
`
`3.1.1. Gel permeation chromatography
`The distributions calculated from the GPC chromatograms of
`the iron preparations show homogenous distributions with the
`
`Pharmacosmos, Exh. 1029, p. 3
`
`

`
`M.R. Jahn et al. / European Journal of Pharmaceutics and Biopharmaceutics 78 (2011) 480–491
`
`483
`
`exception of ferumoxytol and iron carboxymaltose which show
`additional smaller and larger diameter peaks (Fig. 1). The hydrody-
`namic diameters dh rise in the order iron sucrose < sodium ferric
`gluconate < iron isomaltoside 1000 < LMW iron dextran < iron
`carboxymaltose < ferumoxytol (Table 1). Ferumoxytol was eluted
`near the exclusion volume, indicating that both its diameter and
`molecular weight might be underestimated.
`
`3.1.2. Dynamic light scattering (DLS)
`The hydrodynamic diameter determined with DLS also mea-
`sures the carbohydrate shell of the IV iron agents and therefore
`is larger than iron oxide core diameters determined by TEM or
`XRD. In Fig. 2 narrow volume distributions of the whole particle
`diameters are shown. The medians of the hydrodynamic diameters
`rise from 8.3 to 23.6 nm in the order iron sucrose < sodium ferric
`gluconate < iron isomaltoside 1000 < LMW iron dextran  ferum-
`oxytol < iron carboxymaltose (Table 1). The zeta potentials of the
`iron preparations are shown in Table 2. Without pH adjustment,
`all iron preparations are negatively charged with the exception of
`iron carboxymaltose. The order of particle charges starting with
`the most negative iron preparation is ferumoxytol (43.2 mV) < ir-
`on gluconate  iron sucrose < iron isomaltoside 1000 < iron dex-
`tran < iron carboxymaltose
`(+3.7 mV). Acidification of
`the
`samples increased the zeta potential of iron carboxymaltose and
`decreased the negative zeta potential of all other compounds. At
`a pH value close to the physiological pH, all formulations showed
`a negative zeta potential, though that for iron carboxymaltose
`was much smaller.
`
`3.2. Size and structure of core
`
`3.2.1. Transmission electron microscopy (TEM)
`TEM images of IV iron agents are shown in Fig. 3. Dark, electron
`dense, beadlike structures present the cores of the iron oxide com-
`plexes, surrounded by a less electron dense matrix, which may be
`attributed to a carbohydrate fraction. The medians of the geomet-
`rical diameter of the core rise from 4.1 to 6.2 nm in the order so-
`dium ferric gluconate < iron sucrose < LMW iron dextran < iron
`isomaltoside 1000  ferumoxytol (Table 3). In case of iron carboxy-
`maltose cores tend to cluster and single cores are not definable.
`The median geometrical core diameter of
`these clusters is
`11.7 ± 4.4 nm.
`
`Fig. 1. Weight distribution vs. particle diameter as determined by gel permeation
`chromatography.
`
`Table 1
`Shell /Particle dimensions as determined by gel permeation chromatography (GPC)
`and dynamic light scattering (DLS).
`
`Iron complex
`
`MW (kDa)
`
`Calculated
`shell-Ø (nm)c
`
`Shell-Ø (nm)
`
`Sodium ferric gluconate
`Iron sucrose
`LMW iron dextran
`Iron isomaltoside 1000
`Iron carboxymaltose
`Ferumoxytol
`
`GPC
`
`164.1
`140.1
`165.0
`150.0
`233.1
`275.7
`
`20.3
`19.1
`20.7
`20.5
`23.8
`26.3
`
`DLS
`8.6a
`8.3a
`12.2a
`9.9a
`23.1a
`23.6a
`
`0.244b
`0.192b
`0.149b
`0.182b
`0.07b
`0.143b
`
`a Median-Ø.
`b Polydispersity index.
`c The most frequently found particle diameter in the distribution.
`
`Fig. 2. Volume distribution of the hydrodynamic diameter of intravenous iron preparations as determined by dynamic light scattering (DLS). Conditions: 0.4 mg Fe/ml.
`Vertical lines assign the median diameter.
`
`Pharmacosmos, Exh. 1029, p. 4
`
`

`
`484
`
`M.R. Jahn et al. / European Journal of Pharmaceutics and Biopharmaceutics 78 (2011) 480–491
`
`Table 2
`Zeta potentials f of IV iron polynuclear complexes at different pH values.
`
`Iron gluconate
`
`Iron sucrose
`
`LMW iron dextran
`
`Ferumoxytol
`
`Iron isomalto-side 1000
`
`Iron carboxymaltose
`
`pH
`
`4.35
`7.4
`8.36a
`10.5
`
`f (mV)
`16.50
`29.70
`29.10
`29.60
`
`pH
`
`4.49
`7.43
`11.03a
`
`f (mV)
`14.25
`26.20
`28.15
`
`pH
`
`3.02
`6.4a
`7.31
`11.8
`
`f (mV)
`3.56
`15.30
`17.25
`15.75
`
`pH
`
`3.39
`6.6a
`7.36
`10.4
`
`f (mV)
`11.95
`43.20
`30.55
`34.40
`
`pH
`
`3.3
`6.3a
`7.35
`9.03
`11.5
`
`f (mV)
`3.98
`22.00
`21.05
`28.95
`26.40
`
`pH
`
`3.26
`5.36a
`7.26
`9.54
`
`f (mV)
`
`9.46
`3.68
`8.52
`16.35
`
`a pH in bidistilled sterile-filtrated water, without any pH adjustment.
`
`and 56° 2h could be a hint that also other structures are mixed in
`like akaganeite.
`The X-ray results of iron carboxymaltose indicate the akagane-
`ite structure with same intensities at same angles except for a min-
`or peak instead of a major peak at 12° 2h. LMW iron dextran and
`iron isomaltoside 1000 show a pattern which is similar to akagane-
`ite as well, but conformity is not as good (minor peaks instead of
`major peaks at 12° 2h and 35° 2h, in part missing minor peaks).
`The diffractogram of ferumoxytol, which is used as IV iron agent
`and contrast agent in magnetic resonance imaging as well, is close
`to pattern of magnetite and maghemite. Sharp peaks in the diffrac-
`togram belong to crystalline mannitol in the formulation.
`
`3.3. Ferrous iron content
`
`3.3.1. Mössbauer spectroscopy
`The mössbauer spectrum of iron isomaltoside 1000 shows a
`doublet with an isomer shift d = 0.44 mm/s and a quadrupole split-
`ting EQ = 0.78 mm/s (Fig. 5). Both parameters are characteristic for
`iron in the ferric state. There is no indication of iron in the ferrous
`state as characteristic isomer shifts and splittings are absent.
`
`3.4. Dialysable iron content
`
`3.4.1. Dialysis
`The results of the determination of the dialyzable ‘‘free’’ iron
`content are shown in Table 3. It appears that iron isomaltoside
`1000, iron carboxymaltose, and ferumoxytol yield very low free
`iron contents smaller than 0.002% of the total iron content. This
`was independent of the liquid used for the dilution and dialysis
`(water versus sodium chloride solution). Iron dextran yielded free
`iron contents of 0.1% and 0.2% in water and sodium chloride solu-
`tion, respectively. The highest free iron content was observed for
`sodium ferric gluconate yielding more than 1% in sodium chloride
`dilutions. However, the free iron content in the iron sucrose prep-
`aration (0.067% in NaCl and 0.057% in water) was lower than ex-
`pected. The experiment without pH adjustment showed that only
`iron carboxymaltose was affected by pH. As depicted in Fig. 6 the
`content of free iron in iron carboxymaltose increases from below
`the detection limit (<0.002%) at pH 7.5–0.262% when the experi-
`ment is conducted in nonbuffered 0.9% NaCl.
`
`3.5. Labile iron
`
`3.5.1. Acid soluble iron
`In acidic solution, FeOOH is dissociated: FeOOH þ 3HCl !
`Fe3þ þ 3Cl þ 2H2O. In this study, iron formulations [FeOOH]mLn
`with different carbohydrate ligands L were decomposed similarly:
`½FeOOHŠmLn þ 3mHCl ! mFe3þ þ 3mCl þ 2mH2O þ nL.
`As
`the
`molar extinction coefficient of
`the complex at 287.3 nm
`½FeOOHŠmLn  3000 M1 cm1)
`(e287:3nm
`is substantially higher
`than the
`½FeOOHŠmLn  580 M1 cm1) or
`extinction coefficient of Fe3+
`(e287:3nm
`carbohydrate (negligible), the decreasing FeOOH concentration is
`approximately proportional to the measured absorbance.
`
`Fig. 3. Transmission electron microscopy images of intravenous iron preparations.
`Conditions: 1 mg Fe/ml.
`
`3.2.2. X-ray diffraction (XRD)
`The particles mean diameters d of the cores were determined
`using the Scherrer equation and are presented in Table 3. The mean
`diameters of single core complexes are in the range of 3.3–6.4 nm
`and appear in accordance with diameters measured by TEM.
`In Fig. 4, X-ray diffractograms of IV iron agents (upper part of
`figure) are compared with diffraction data of standard iron oxides
`from the ICDD (lower part of figure, International Centre for Dif-
`fraction Data). Peaks belonging to the carbohydrate fraction are
`marked with arrows. With the exception of ferumoxytol the IV iron
`agents show broad regions of high intensities at in part similar an-
`gle values of diffraction with similar intensities.
`The patterns of iron sucrose and sodium ferric gluconate show a
`structure similar to 2-line ferrihydrite as there are just two major
`iron oxide peaks at 36° 2h and 62° 2h. Two others at 14° 2h and 22°
`2h belong to amorphous sucrose [27]. Small reflections at 40° 2h
`
`Pharmacosmos, Exh. 1029, p. 5
`
`

`
`M.R. Jahn et al. / European Journal of Pharmaceutics and Biopharmaceutics 78 (2011) 480–491
`
`485
`
`Table 3
`Core dimensions as determined by transmission electron microscopy (TEM) and X-ray diffraction (XRD), acidic hydrolysis stability and dialysable iron with and without pH
`adjustment.
`
`Iron complex
`
`Sodium ferric gluconate
`Iron sucrose
`LMW iron dextran
`Iron isomaltoside 1000
`Iron carboxymaltose
`Ferumoxytol
`
`Core-Ø (nm)
`
`TEM
`4.1a ± 1.7
`5.0a ± 0.8
`5.6a ± 1.2
`6.3a ± 1.2
`11.7a,b ± 44
`6.2a ± 1.4
`
`t0.5 (h)
`
`XRD
`
`Acidic hydrolysis
`
`3.4
`3.3
`4.4
`4.2
`4.3
`6.4
`
`4.0 ± 0.1
`4.9 ± 0.1
`21.0 ± 1.2
`25.2 ± 1.2
`25.6 ± 1.6
`62.4 ± 0.4
`
`Dialysable irone (%)
`WFIf
`
`0.789 ± 0.048
`0.057d
`0.100 ± 0.0096
`<0.002c
`<0.002c
`<0.002c
`
`NaClf
`1.338d
`0.067d
`0.207 ± 0.0071
`<0.002c
`<0.002c
`<0.002c
`
`NaClg
`
`0.172h ± 0.0048
`0.014h ± 0.0029
`0.2621h ± 0
`0.005h ± 0.0047
`
`a Median-Ø.
`b Median-Ø of an agglomeration of several cores. Single cores are not definable.
`c Detection limit.
`d Only one sample.
`e Calculated from ICP-MS measurements.
`f Adjustment to pH 7.5.
`g Without pH adjustment.
`h Product approved for high dose administration.
`
`Fig. 5. Mössbauer spectrum of MonoFer as powder measured at 150 K. Measuring
`points are fitted with a line of Lorentz shape.
`Isomer shift d = 0.44 mm/s,
`quadrupole splitting EQ = 0.78 mm/s, line widths U = 0.69 mm/s.
`
`administered dose of 200 mg and 500 mg, respectively. The prod-
`ucts which show the highest fraction of labile, FerrozineÒ-detect-
`able iron are, by far, iron gluconate and iron sucrose (3.2 ± 0.4%
`for iron gluconate, 3.5 ± 0.2% for iron sucrose at a 200 mg dose,
`respectively). For the different compounds, the fraction of labile
`
`Fig. 4. X-ray spectra of intravenous iron preparations. At the bottom, the spectra
`are compared with diffraction data of standard iron oxides from the ICDD
`(International Centre for Diffraction Data). Thick reflexion lines of standard iron
`oxides: Intensity 70–100%. Thin reflexion lines of standard iron oxides: 30–69%
`intensity. Peaks belonging to carbohydrate fraction are assigned by arrows.
`
`The rate of hydrolysis is a measure of the relative stability of the
`FeOOH entity. In Fig. 7, it can be seen that the fraction of FeOOH
`remaining decreases with time, and in Table 3, the half-times of
`FeOOH decomposition of the various iron preparations are com-
`pared. The complex stability is increasing in the order of iron glu-
`conate < iron sucrose  iron dextran < iron carboxymaltose  iron
`isomaltoside 1000  ferumoxytol. There is some indication that
`the rate of degradation relates to the surface area of the iron com-
`plex formulation and decreasing with increasing particle size
`(Fig. 8).
`
`3.5.2. FerrozineÒ-detectable labile iron in human serum
`The results of the determination on detectable labile iron with
`the FerrozineÒ-method are shown in Fig. 9. The amount of the
`labile iron, measured by this test was nearly equivalent to the
`
`Fig. 6. Comparative free iron content in high dose IV iron products. The detection
`limit was 0.002%. Red bars indicate free iron content following adjustment of the
`diluted preparation to pH 7; blue bars indicate results obtained without pH
`adjustment. Star indicates concentrations below detection limit. SD’s are listed in
`Table 3.
`
`Pharmacosmos, Exh. 1029, p. 6
`
`

`
`486
`
`M.R. Jahn et al. / European Journal of Pharmaceutics and Biopharmaceutics 78 (2011) 480–491
`
`sodium ferric gluconate
`iron sucrose
`LMW iron dextran
`iron isomaltoside
`iron carboxymaltose
`ferumoxytol
`20
`
`30
`
`Time [h]
`
`40
`
`50
`
`0
`
`10
`
`0.00
`
`-0.25
`
`-0.50
`
`-0.75
`
`-1.00
`
`-1.25
`
`-1.50
`
`-1.75
`
`-2.00
`
`ln(fraction FeOOH)
`
`Fig. 7. Acid soluble iron. Concentration: 10 lg Fe/ml, 0.2375 M HCl. At t = 0 min, the
`fraction of FeOOH is 1 according to 100% not hydrolyzed FeOOH. Each point
`represents the average of three measurements; error bars are sometimes smaller
`than symbols. Data were fitted with a second degree polynomial (R2 > 0.990). The
`solid line labels the half-time. SD’s are listed in Table 3. For further details refer to
`methods.
`
`Fig. 9. Comparative labile iron pools of parenteral iron products. Upper diagram:
`Concentration of the FerrozineÒ-detectable labile iron pool in lg/ml. The bars
`represent the average of at least four measurements. Lower diagram: FerrozineÒ-
`detectable labile iron in percentage of the total used dose. For each measurement,
`the iron complex was incubated in human serum for 10 and 45 min, respectively.
`Thereafter, the FerrozineÒ reaction was performed and in each case an intercept of a
`second degree polynomial regression function of the absorption versus time curve
`with the ordinate was calculated. These intercepts were extrapolated to an
`incubation time of t = 0 by linear regression, yielding the labile iron pool in serum
`for each intravenous iron product.
`
`200 MHz of isomaltoside 1000 confirmed the above-mentioned
`conclusions as shown in Fig. 10a. Data of 13C NMR measurements
`for the iron isomaltoside 1000 complex as prepared described in
`patent [26] showed line broadening of signals as seen in Fig. 10b.
`The spectrum demonstrates a significant line broadening of the
`carbon signals carbon C-1, C-5, and C-6 and a smaller line broaden-
`ing of C-3 and C-2 al

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