throbber
+
`
`+
`
`January 1997
`Volume 86, Number 1
`
`REVIEW ARTICLE
`
`Characteristics and Significance of the Amorphous State in Pharmaceutical
`Systems
`
`BRUNO C. HANCOCK*X AND GEORGE ZOGRAFI†
`Received April 26, 1996, from the *Merck Frosst Canada Inc., Pointe Claire-Dorval, Quebec, H9R 4P8, Canada, and †School of Pharmacy,
`University of WisconsinsMadison, Madison, WI 53706.
`Final revised manuscript received July 31, 1996.
`Accepted for
`publication August 1, 1996X.
`
`Abstract 0 The amorphous state is critical
`in determining the solid-
`state physical and chemical properties of many pharmaceutical dosage
`forms. This review describes the characteristics of the amorphous state
`and some of the most common methods that can be used to measure
`them. Examples of pharmaceutical situations where the presence of the
`amorphous state plays an important role are presented. The application
`of our current knowledge to pharmaceutical
`formulation problems is
`illustrated, and some strategies for working with amorphous character in
`pharmaceutical systems are provided.
`
`Introduction
`During the final stage of developing a synthetic procedure
`for a new drug entity, a great deal of emphasis is placed on
`obtaining material of high purity, and reproducibility in terms
`of its physical, chemical, and biological properties. Every
`effort is made to ensure a high degree of crystallinity, wherein
`the molecules have regular and well-defined molecular pack-
`ing, and emphasis is also placed on whether or not the
`compound can exist in polymorphic or solvated crystal forms.1
`These forms can have different thermodynamic properties
`(e.g., melting temperature, vapor pressure, solubility), and a
`knowledge of their existence is required to anticipate spon-
`taneous changes in the properties of the solid during storage
`and/or handling of the material. It is also possible that upon
`isolation the material will be obtained in a fully or partially
`amorphous state.2 The four most common means by which
`amorphous character is induced in a solid are shown in Figure
`1. These are condensation from the vapor state, supercooling
`of the melt, mechanical activation of a crystalline mass (e.g.,
`during milling), and rapid precipitation from solution (e.g.,
`during freeze-drying or spray drying). Amorphous character
`is common with polymeric molecules used as excipients, and
`
`X Abstract published in Advance ACS Abstracts, October 1, 1996.
`
`© 1997, American Chemical Society and
`American Pharmaceutical Association
`
`S0022-3549(96)00189-X CCC: $14.00
`
`Figure 1sSchematic diagram of the most common ways in which amorphous
`character is induced in a pharmaceutical system.
`
`large peptides and proteins used as therapeutic agents, and
`it can also occur with small organic and inorganic molecules.
`When a system consists of multiple components, as with
`pharmaceutical formulations, it is possible that amorphous
`solid-state solutions can form analogous to liquid solutions.
`Water vapor can also be absorbed by an amorphous solid to
`form an amorphous solid solution.
`The three-dimensional
`long-range order that normally
`exists in a crystalline material does not exist in the amorphous
`state, and the position of molecules relative to one another is
`more random as in the liquid state. Typically amorphous
`solids exhibit short-range order over a few molecular dimen-
`sions and have physical properties quite different from those
`of their corresponding crystalline states.
`In Figure 2 we
`schematically plot the enthalpy (H) or specific volume (V) of
`a solid substance as a function of its temperature. For a
`crystalline material at very low temperatures we see a small
`increase in enthalpy and volume with respect to temperature,
`indicative of a certain heat capacity (Cp) and thermal expan-
`sion coefficient (R). There is a discontinuity in both H and V
`at the melting temperature (Tm) representing the first-order
`phase transition to the liquid state. Upon rapid cooling of
`the melt the values of H and V may follow the equilibrium
`line for the liquid beyond the melting temperature into a
`Argentum EX1028
`Journal of Pharmaceutical Sciences / 1
`Vol. 86, No. 1, January 1997
`
`Page 1
`
`

`

`+
`
`+
`
`this state by considering its rate and extent of molecular
`motions. The average time scale of molecular motions within
`a supercooled liquid is usually less than 100 s, the viscosity
`is typically between 10-3 and 1012 Pa(cid:226)s (Figure 3), and both
`properties are strongly temperature dependent.6-10 Cooling
`the supercooled liquid even further appears to reduce the
`molecular mobility of the material to a point at which the
`material is kinetically unable to attain equilibrium in the time
`scale of the measurement as it loses its thermal energy,
`resulting in a change in the temperature dependence of the
`enthalpy and volume. The temperature at which this occurs
`is the experimentally observed glass transition temperature
`(Tg). Below Tg the material is “kinetically frozen” into a
`thermodynamically unstable glassy state with respect to both
`the equilibrium liquid and the crystalline phase, and any
`further reduction in temperature has only a small effect upon
`its structure. Molecular motions in glasses typically occur
`over a period in excess of 100 s, and viscosities are usually
`greater than 1012 Pa(cid:226)s.6-10 Many of the physical properties
`of glassy amorphous materials (e.g., thermal expansion coef-
`ficient) are different from those of the corresponding super-
`cooled liquid above Tg.
`The molecular processes which contribute to the glass
`transition are currently the subject of intensive research and
`debate. Whether the changes in thermodynamic properties
`(e.g., specific heat, volume) that are seen during cooling (or
`reheating) are due to a real thermodynamic phase transition
`or are of purely kinetic origin is a controversial issue, and no
`theory has yet been proposed which accounts for all the
`observed experimental features. Several excellent reviews
`which describe the current thinking in this field have been
`published.6-8,10,11 Models based on statistical mechanical or
`free volume theories are the simplest and most widely
`invoked. Polymer scientists, metallurgists, ceramists, etc.
`each have their preferred approaches with specific advantages
`for the materials and processes with which they are working.
`From Figure 2 it can be seen that the glass transition can be
`considered to be a thermodynamic requirement for a super-
`cooled liquid since without such a transition the amorphous
`material would attain a lower enthalpy than the crystalline
`state at some critical temperature and would eventually attain
`a negative enthalpy. This critical temperature is known as
`the Kauzmann temperature (TK) and is thought to mark the
`lower limit of the experimental glass transition (Tg) and to
`be the point at which the configurational entropy of the system
`reaches zero.9,10 Experimental studies of the glass transition
`are complicated by the existence of many different modes of
`molecular motion in most systems (e.g., rotational or trans-
`lational), changes in the scale and type of motions with
`temperature, and cooperativity or coupling of molecular
`motions. One can only say for certain that at Tg the mean
`molecular relaxation time ((cid:244)) associated with the predominant
`molecular motions is about 100 s and that Tg can be expected
`to vary with experimental heating and cooling rates, sample
`molecular mass,12,13 sample history, sample geometry,14,15 and
`sample purity.16 The experimental glass transition temper-
`ature is also influenced by the choice of technique used to
`measure it because of the varying sensitivities of available
`techniques to different types and speeds of molecular motions.
`The temperature dependence of molecular motions directly
`determines many important physical properties of amorphous
`materials,
`including the location of the glass transition
`temperature and the ease of glass formation. This tempera-
`ture dependence is most frequently described using the
`empirical Vogel-Tammann-Fulcher (VTF) equation:7,8,10
`(cid:244) ) (cid:244)0 exp(DT0/T-T0)
`
`(1)
`
`where (cid:244) is the mean molecular relaxation time, T is the
`
`Figure 2sSchematic depiction of
`temperature.
`
`the variation of enthalpy (or volume) with
`
`“supercooled liquid” region. On cooling further a change in
`slope is usually seen at a characteristic temperature known
`as the glass transition temperature (Tg). At Tg the properties
`of the glassy material deviate from those of the equilibrium
`supercooled liquid to give a nonequilibrium state having even
`higher H and V than the supercooled liquid. As a result of
`its higher internal energy (e.g., (cid:25)25 kJ(cid:226)mol-1 for cephalospor-
`ins3) the amorphous state should have enhanced thermody-
`namic properties relative to the crystalline state (e.g., solu-
`bility,4 vapor pressure) and greater molecular motion. We
`would also expect amorphous systems to exhibit greater
`chemical reactivity3 and to show some tendency to spontane-
`ously crystallize, possibly at different rates below and above
`Tg.5 From a pharmaceutical perspective we have an interest-
`ing situation. The high internal energy and specific volume
`of the amorphous state relative to the crystalline state can
`lead to enhanced dissolution and bioavailability,4 but can also
`create the possibility that during processing or storage the
`amorphous state may spontaneously convert back to the
`crystalline state.5
`In considering the importance of the amorphous state in
`pharmaceutical systems we must direct our attention to two
`main situations. In the first, a material may exist intrinsically
`in the amorphous state or it may be purposefully rendered
`amorphous and we would like to take advantage of its unique
`physical chemical properties. Under these circumstances we
`usually want to develop strategies to prevent physical and
`chemical instability of the amorphous sample. In the second
`case, we may be dealing with a crystalline material that has
`been inadvertently rendered amorphous during processing.
`This type of amorphous character usually exists predomi-
`nately at surfaces at levels not easily detected and has the
`potential to produce unwanted changes in the physical and
`chemical properties of the system. In this situation we usually
`want to process the system so that the amorphous portions
`of the solid are converted back to the most thermodynamically
`stable crystalline state.
`
`Definition and Description of the Amorphous State
`The rapid cooling of a liquid below its melting point (Tm)
`may lead to an amorphous state with the structural charac-
`teristics of a liquid, but with a much greater viscosity (Figures
`2 & 3). The enthalpy and volume changes immediately below
`Tm exhibit no discontinuity with those observed above Tm, so
`we consider this amorphous state to be an equilibrium
`“supercooled” liquid. This amorphous state is also called the
`“rubbery state” because of the macroscopic properties of
`amorphous solids in this region. We can further characterize
`
`2 / Journal of Pharmaceutical Sciences
`Vol. 86, No. 1, January 1997
`
`Page 2
`
`

`

`+
`
`+
`
`Figure 3sMolecular mobility (or viscosity) of amorphous materials as a function of normalized temperature above Tg.7,8 Reprinted with permission from ref 8. Copyright
`1995 American Association for the Advancement of Science.
`
`temperature, and (cid:244)0, D, and T0 are constants. The value of
`T0 in the VTF equation is believed to correspond to the
`theoretical Kauzmann temperature (TK), and (cid:244)0 can be related
`to the relaxation time constant of the unrestricted material.7,8
`When T0 is 0, the familiar Arrhenius equation is obtained,
`and D is directly proportional to the activation energy for
`molecular motion. When T0 is greater than 0, there is a
`temperature dependent apparent activation energy. The
`Williams-Landel-Ferry (WLF) equation17 describing the
`temperature dependence of viscosity (Ł) in polymers above Tg
`is a special case of the VTF equation:
`Ł ) Łg exp{C1(T - Tg)/(C2 + (T - Tg))}
`
`(2)
`
`where Łg is the mean viscosity at Tg and C1 and C2 are
`constants. This equation can be derived from first principles
`based on polymer free volume theories. The constants C1 and
`C2 are found to be quite universal for a range of polymers17
`and are equivalent to DT0/(Tg - T0) and (Tg - T0), respectively,
`in the VTF equation. The WLF equation has been shown to
`fit viscosity data for several small organic molecules using
`the universal constants,18-20 making it useful for predicting
`the relaxation behavior or molecular mobility of amorphous
`pharmaceutical solids. However, it is important to recognize
`that this is not always the case and such predictions cannot
`always be assumed to be accurate.
`Depending upon the magnitude and temperature depen-
`dence of the apparent activation energy for molecular motions
`near and above Tg in supercooled liquids, it is possible to
`classify them as either “strong” or “fragile” amorphous systems
`(Figure 3).7,8 A strong liquid typically exhibits Arrhenius-like
`changes in its molecular mobility with temperature and a
`relatively small change in heat capacity at Tg. Proteins are
`good examples of strong glass formers, with their changes in
`heat capacity at Tg often being so small that they cannot be
`detected using standard calorimetry techniques.21 A fragile
`
`supercooled liquid has a much stronger temperature depen-
`dence of molecular mobility near Tg and a relatively large
`change in heat capacity at Tg and will typically consist of
`nondirectionally, noncovalently bonded molecules (e.g., etha-
`nol). The constant D in the VTF equation is an indicator of
`fragility, with low values (<10) corresponding to very fragile
`glass formers and high values (>100) indicating strong glass-
`forming tendencies. The value of T0 in the VTF equation is
`also linked to the fragility of the system with (Tg - T0) > 50
`typical of strong glass formers and (Tg - T0) < 50 usual for
`fragile materials. A simple graphical means of ranking
`materials in terms of their strength/fragility is to plot the
`molecular mobility (or viscosity) as a function of the temper-
`ature normalized to the experimental glass transition tem-
`perature (e.g., Figure 3).7,8 A “rule of thumb” for determining
`fragility without relaxation time data has also been proposed
`based on the relative magnitudes of the melting and glass
`transition temperatures: strong, Tm/Tg (in K) >1.5; fragile,
`Tm/Tg (in K) < 1.5)7, 8 (Table 1). (See Note Added in Proof.)
`The extent of departure of a glass’s properties from equi-
`librium is determined by its formation conditions, so we can
`presume the existence of multiple metastable glasses below
`Tg (Figure 2),2,3 and even polyamorphic glasses that convert
`via first-order transitions.22-24 As a result of this, the tem-
`perature dependence of molecular motions below the glass
`transition temperature is highly dependent upon the condi-
`tions under which the glass was formed.12 This temperature
`dependence is generally less extreme than above Tg and more
`linear, with some authors proposing an Arrhenius-like rela-
`tionship. That molecular motions do occur below Tg is
`unquestionable, and the consequences of the relaxation or
`“aging” of glassy materials have been widely reported. For
`example, Guo et al.25 described effects upon the film-coat water
`permeability and dissolution rate of film coated tablets, and
`Byron and Dalby26 studied the effects of aging on the perme-
`ability of poly(vinyl alcohol) films to a model water soluble
`
`Journal of Pharmaceutical Sciences / 3
`Vol. 86, No. 1, January 1997
`
`Page 3
`
`

`

`+
`
`+
`
`Table 1sMeasured Physical Properties of Some Amorphous Pharmaceutical Materialsa
`
`Material
`
`Mw
`
`Tm (K)
`
`Tg (K)
`
`Tm/Tg
`
`Famorph (kg(cid:226)m-3)
`Fcrystal (kg(cid:226)m-3)
`¢Cp (J(cid:226)g-1(cid:226)K-1)
`1.32
`1.38
`0.466
`1.37
`320
`438
`358
`Indomethacin
`1.43
`1.59
`0.544
`1.30
`348
`453
`342
`Sucrose
`1.48
`1.60
`0.472
`1.27
`383
`486
`342
`Lactose (anhydrous)
`1.49
`1.58
`0.534
`1.24
`385
`476
`342
`Trehalose (anhydrous)
`(cid:25)5 · 105
`s
`s
`s
`Dextran
`0.92
`0.400
`498
`(cid:25)1 · 106
`s
`s
`s
`Poly(vinylpyrrolidone)
`1.25
`0.260
`458
`(cid:25)0.95
`Waterb
`<0.95
`0.100
`2.01
`136
`273
`18
`aMw ) molecular mass; Tg ) glass transition temperature; Tm ) melting temperature; ¢Cp ) heat capacity change at Tg; F ) density. b Reference 134.
`
`drug. The effects of aging are often detrimental, but they can
`also be used to improve a product’s performance with a
`deliberate “annealing” process. This strategy is particularly
`useful when small amounts of amorphous character have been
`unintentionally introduced into a system by high-energy
`processing (see later).27,28 The time scale of molecular motions
`in a glass is much longer than above Tg ((cid:244) . 100 s) and
`requires different experimental techniques for its study. In
`almost all cases the molecular relaxation processes that occur
`in glasses follow a nonexponential function. This nonexpo-
`nentiality has been widely studied and modeled29 and appears
`to be the result of a heterogeneous microstructure within
`glasses which leads to a distribution of types and rates of
`molecular motion under any given time and temperature
`conditions. The reader is referred to some excellent reviews
`for detailed information on the application of these models to
`glassy systems.12,29 The empirical Kohlrausch-Williams-
`Watts (KWW) stretched exponential function is most often
`used to describe the distribution of molecular motions:10
`(cid:30)(t) ) exp{-(t/(cid:244))(cid:226)}
`
`(3)
`
`where (cid:30)(t) is the extent of relaxation at time t, (cid:244) is the mean
`molecular relaxation time, and (cid:226) is a constant. A (cid:226) value of
`unity corresponds to a single relaxation time with exponential
`behavior. The smaller the value of (cid:226), the more the distribu-
`tion of molecular motions deviates from a single exponential.
`(cid:226) has been shown to correspond to the strength/fragility of
`the material above Tg, but no similar relationship has yet been
`established below Tg. By fitting data to the KWW function it
`is possible to determine the mean molecular relaxation time
`((cid:244)) and (cid:226) for any well-defined glass.30 A general means of
`ranking glasses in terms of the temperature dependence of
`molecular motions, similar to Angell’s strong/fragile classifica-
`tion system above Tg, would be of great use to pharmaceutical
`materials scientists but has not yet been developed because
`of the greater complexities of the glassy state.
`Perhaps the most important question relating to amorphous
`pharmaceutical systems is, At what temperature do the
`molecular motions responsible for physical and chemical
`instabilities cease to become likely over the lifetime of that
`particular system?30 It has been suggested that this lower
`temperature limit might correspond to the Kauzmann tem-
`perature (TK). Although this appears to be the case for some
`systems, there also appears to be an influence from the
`strength/fragility of the system, and also from whether or not
`the molecular motions that are responsible for the glass
`transition and any instabilities are identical.30 Mean molec-
`ular relaxation times have been reported for several pharma-
`ceutical glass formers as a function of temperature following
`enthalpy relaxation and thermomechanical relaxation experi-
`ments, and the temperature of negligible molecular mobility
`during a 3-year shelf life varied according to (i) the method
`used to assess the molecular motions and (ii) the identity of
`the glass former.30 As yet there is no reliable means of
`predicting the temperature of negligible molecular mobility
`in amorphous solids, and thus a conservative approach is
`
`4 / Journal of Pharmaceutical Sciences
`Vol. 86, No. 1, January 1997
`
`required when defining storage and processing conditions for
`amorphous pharmaceutical systems (see later).
`The behavior of amorphous systems as defined in Figure 2
`is dependent upon the assumption of constant pressure and
`composition. Pressure effects upon amorphous materials have
`not been widely studied but are likely to be significant with
`effects on molecular packing potentially modifying the glass
`transition temperature, the thermal expansion behavior, and
`the strength/fragility of a supercooled liquid.10,31,32 From a
`practical perspective the glass transition temperature of a
`system containing volatile components may only be experi-
`mentally accessible at elevated pressures. For example, the
`widespread and significant plasticizing effects of sorbed water
`vapor in high-Tg amorphous polymers have only recently been
`fully realized because of advances in sample-handling methods
`which allow samples of varying water content to be sealed at
`ambient temperature and then heated through Tg without loss
`of their sorbed water vapor.33 The properties of a glassy
`amorphous solution prepared by lyophilization are also likely
`to be significantly different from those of the same system
`prepared at ambient pressure since the reduced pressure
`within a lyophilization chamber will affect the structure of
`the amorphous cake that is formed and also the composition
`of the solution through the primary and secondary drying
`processes. Angell et al.34 have noted that for aqueous solutions
`the fragility of the supercooled solution is dependent upon the
`solute concentration in the solution. From the limited data
`available it can be concluded that some supercooled aqueous
`solutions become stronger as they become more dilute (e.g.,
`sugars), whereas others become more fragile (e.g., electrolytes,
`salts). The type of behavior observed appears to be linked to
`the extent of hydrogen bonding in the aqueous solution. The
`fragility of such mixed systems may also be related to the
`ideality of their mixing behavior. Simple mixing rules have
`been used by many authors35,36 to describe the variation of
`the glass transition temperature with blend composition;
`however, the effects of nonidealities (e.g., immiscibility, mo-
`lecular size differences, specific interactions, etc.) are often
`significant. The simplest and most reliable approach for use
`with amorphous pharmaceutical materials appears to be a
`modified Gordon-Taylor equation35,36 which is based on free
`volume theories with some simplifying assumptions. For
`simple two-component mixtures,
`Tgmix ) (w1Tg1 + Kw2Tg2 )/(w1 + Kw2)
`
`(4)
`
`where Tg is the glass transition temperature, w1 and w2 are
`the weight fractions of components 1 and 2, and K can be
`calculated from the densities (F) and glass transition temper-
`atures (Tg) of the components:
`K ) (Tg1F1)/(Tg2F2)
`
`(5)
`
`Similar equations can be readily derived for mixtures of more
`than two components. A perfecty miscible system will display
`a single sharp glass transition event. Immiscibility, incom-
`patibility, or nonideality is often indicated by a poor fit to the
`
`Page 4
`
`

`

`+
`
`+
`
`theoretical equation, the appearance of more than one Tg, or
`“broadening” of the glass transition event. Deviations from
`ideal behavior can also be identified and their most likely
`causes assessed using the graphical approach of Schneider
`and co-workers.36,37 Deviations usually occur over discrete
`composition ranges and often can be explained in terms of
`molecular size effects and the disappearing free volume of the
`high-Tg component at lower temperatures and composi-
`tions.38,39 Such an approach is analogous to percolation
`theories and has considerable potential for describing mixed
`amorphous systems. Simple solution theories also can be used
`to describe such systems and to provide a qualitative under-
`standing of the important factors regulating the glass transi-
`tion in pharmaceutical systems. For example, it is likely that,
`when a macromolecule is mixed in small amounts with an
`amorphous small molecule, it will introduce a considerable
`excess free volume to the system because of its much larger
`molecular size. In this situation the glass transition temper-
`ature of the mixture probably will not be elevated as much
`as predicted by theory. The addition of low levels of a small
`molecule to an amorphous macromolecular system probably
`will be much less disruptive. Both materials will make near
`ideal contributions to the overall free volume of the mixture,
`and in this instance the predictions of the mixing equations
`are likely to be quite accurate for at least the first 50 K change
`in Tg. This is very important since the presence of very low
`levels of low molecular weight contaminants or additives
`(including water vapor) is predicted and observed to have
`significant plasticizing effects on pharmaceutical glasses,36
`whereas the addition of low levels of high molecular mass
`additives often has minimal antiplasticizing effect.39 It should
`be noted that the concept of a critical additive composition
`(Wg) at which a glassy macromolecular material is sufficiently
`plasticized by a low molecular weight penetrant that it
`transforms to a rubbery amorphous solid under ambient
`conditions has been described by several authors.33,40
`Pharmaceutical solids rarely exist as 100% crystalline or
`100% amorphous phases so it is necessary at this point to
`consider how partially crystalline or amorphous systems are
`likely to behave. The coexistence of two thermodynamically
`different states of a material will probably result in (i)
`significant and measurable structural heterogeneities and (ii)
`batch to batch variations in physical properties. The presence
`of one phase in another can act as a focal point for spontaneous
`phase transitions such as crystallization.28,41,42 In addition,
`as each phase is intimately dispersed in the other, there may
`not be complete independence of their behavior. For example,
`the dispersion of crystalline drug in an amorphous carrier has
`been reported to alter the observed glass transition temper-
`ature of the amorphous phase.43 For macromolecules there
`may even be molecules which are part of both crystalline and
`amorphous domains physically linking the two regions to-
`gether. Partially ordered systems have traditionally been
`described using either “one-state” or “two-state” models.4,28,41,42
`In the two-state model, domains of material are assumed to
`be either 100% amorphous or 100% crystalline and they
`coexist side by side in a molecular mixture. This type of
`system can be simulated to some extent by making physical
`mixtures of reference samples of crystalline and amorphous
`materials.3 The one-state model consists of domains which
`are truly partially crystalline and in which the molecules have
`formed a semiordered structure as a result of being restricted
`in their motion during crystallization, or following the disrup-
`tion of a more perfect crystalline state. The one-state model
`seems intuitively more likely than the two-state model but
`raises many questions which cannot be readily answered by
`studying mixtures of the reference crystalline and amorphous
`materials. In metallic systems there is also a state known as
`the “nanocrystalline phase” which has properties intermediate
`
`Figure 4sX-ray powder diffraction patterns for amorphous (bottom) and crystalline
`(top) lactose.
`
`to those of the amorphous and crystalline states,44 and the
`concept of “glassy” or “plastic” crystals has recently been
`described.45 Clearly the ability to distinguish between crys-
`talline and amorphous states of a material and to be able to
`quantify one phase in the presence of the other is critical to
`the successful design and production of amorphous pharma-
`ceutical systems.
`
`Characterization of the Amorphous State
`Upon passing into the supercooled liquid state or through
`the glass to rubber transition it is possible to observe changes
`in a multitude of material physical properties including
`density, viscosity, heat capacity, X-ray diffraction, and diffu-
`sion behavior. Techniques which measure these properties
`(directly or indirectly) can be used to detect the presence of
`an amorphous material (glass or rubber), and some of these
`methods are sensitive enough to allow quantification of the
`amount of molecular order or disorder (amorphous content)
`in a partially crystalline system.
`As there is no long-range three-dimensional molecular order
`associated with the amorphous state, the diffraction of
`electromagnetic radiation (e.g., X-rays) is irregular compared
`to that in the crystalline state (Figure 4). Diffraction tech-
`niques are perhaps the most definitive method of detecting
`and quantifying molecular order in any system, and conven-
`tional, wide-angle and small-angle diffraction techniques have
`all been used to study order in systems of pharmaceutical
`relevance.3,5,41 The specificity and accurate quantitative
`nature of these nondestructive techniques make them first
`line choices for studying partially crystalline pharmaceutical
`materials. Conventional X-ray powder diffraction measure-
`ments can be used to quantify non-crystalline material down
`to levels of about 5%41 and with temperature and environ-
`mental control can also be used to follow the kinetics of phase
`transformations, or to quantify the presence of a crystalline
`drug in an amorphous excipient matrix.46 Small-angle X-ray
`measurements have been used to study subtle structural
`(density) changes in polymers in the glassy state upon
`annealing,47 and neutron scattering is gaining wider use in
`the characterization of short-range two-dimensional order in
`amorphous materials.48 It should be remembered that dif-
`fraction techniques only “see” molecular order, and thus
`disorder is only implied.
`The irregular arrangement of molecules in the amorphous
`state usually causes them to be spaced further than in a
`crystal so that the specific volume is greater and the density
`lower than that of the crystal, and we say that there is a
`greater “free volume” (Figure 2). Highly accurate measure-
`
`Journal of Pharmaceutical Sciences / 5
`Vol. 86, No. 1, January 1997
`
`Page 5
`
`

`

`+
`
`+
`
`ments of volume or density are difficult, and the magnitudes
`of differences involved are fairly small (Table 1), so such
`determinations are not a method of choice for characterizing
`amorphous pharmaceutical systems. The use of gas displace-
`ment pycnometry for quantifying the amorphous content of
`partially crystalline pharmaceutical systems has been de-
`scribed by Saleki-Gerhardt et al.,41 and the accuracy achieved
`was about (10%. Liquid displacement pycnometry has also
`been used to determine the crystallinity of several starch
`samples.49 This approach has proved useful for quantifying
`low levels of disorder in crystalline pharmaceutical samples.50
`Precise dilatometric techniques are widely used for the study
`of amorphous polymers,51 and these techniques could be useful
`for pharmaceutical systems; however, the methods are time-
`consuming and quite difficult to perform and so would not be
`suitable for routine use.
`The most characteristic property of the amorphous state is
`its viscosity (approximately <1012 Pa(cid:226)s above Tg and >1012
`Pa(cid:226)s below Tg). The greater free volume and molecular
`disorder of an amorphous material compared to its crystalline
`counterpart result in the mechanical properties of amorphous
`systems (viscosity, elastic modulus) being much more like
`those of a liquid than those of a solid. In some supercooled
`liquids a breakdown of the Stokes-Einstein relationship
`between the viscosity and the rate of diffusion (or molecular
`mobility) has been observed.52 It is thought that the break-
`down of this fundamental relationship may be due to the
`decoupling of certain modes of molecular motion in the
`amorphous state.52,53 Methods which are used to measure the
`viscosity of amorphous materials are quite specialized and
`include the bending of rods or curved fibers below Tg
`54 and
`the torsion pendulum and falling-sphere methods above Tg.19
`Experimental difficulties are often quite significant, and thus
`such measurements are usually feasible only in a specially
`equipped research laboratory.
`Diffusion-controlled processes (such as gas transport, self-
`diffusion, crystallization, and some chemical reactions) which
`are all closely linked to the viscosity of the amorphous matrix
`have been the subject of many studies of glassy and rubbery
`amorphous materials.23 As a consequence of the greater free
`volume in amorphous materials, diffusive transport processes
`are usually significantly more rapid and isotropic than in
`crystals. Often different diffusion rates are observed above
`and below Tg, with significantly faster diffusion occurring
`above Tg. This phenomenon can be used to identify the glass
`transition temperature in some amorphous systems. The
`sensitivity of transport processes to small fractions of crystal-
`line order has yet to be clearly defined, although it appears
`that ordered domains can slow down the rate of diffusion.
`There is an intuitive connection between the viscosity of an
`amorphous system and the rate of diffusion wit

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