throbber
Why molecules move along a temperature gradient
`
`Stefan Duhr and Dieter Braun*
`
`Chair for Appli~d Physic;, Ludwig MaJ!imilians Universitat, Amalienstr<me 5•1, 80799 Mtmich, Germany
`
`Edited by Leo P. Kadanoff, University of Chkago, Chicago, IL, and approved Octob<:r 12, 2006 {rec.eived for review M;~y 26, 2006)
`
`Molecules drift along temperature gradients, an effect called
`thermophoresi$, the Soret effect, or thermodiffuslon.ln liquids, it~
`theoretical foundation is the $Ubject of a long-standing debate. By
`using an all-optkai microfiuidic fluore$cem:e method, we present
`experimental results for DNA and polystyrene beads over a large
`range of partide s!a:e~. salt concentrations, and temperatllri1!5. The
`data support a unifying theory based on solvation entropy. Stated
`in !;imple terms, the Soret meffldrmt i!i given by the negative
`solvation entropy, dh1ided by kT, Thil th!lory predicts the thermod(cid:173)
`iffusion of polystyrenil beads and DNA without any free parame·
`tiH$. We as;ume a local thermodynamic equilibrium of the •oivent
`molecules around the molecule. This as!lumptlon is fulfilled for
`moderate temperature gradients below a fluctuation criterion. For
`both DNA and polystyrene beads, thermophoretk motion d-sange$
`sign at lower temperatures. This thermophilidty toward lower
`temperatures is attributed to an il"lcrea$lng positive sntropy of
`hydration, whereas the generally dominating thermophobkity is
`explained by the negative entropy of ionk !ihielding. The under·
`standing of thermodiffusion set!i the ~tage for detailed probing of
`solvation pmpertie!i of colloids and biomolewles. For exam pi!<!, we
`!iUCCI!5sfu!ly determine the etrectiva charga of DNA and bead> over
`a size range that Is not acces5ibie with electrophoresis.
`
`DNA ! flworescer.ce I micrctluidic ! Soret I th~rmodiffusion
`
`T hermodiffusion has been known for a long time (1), but its
`
`theoretical explanation for mole.cules in liquids is still under
`debate. The search for a theoretical understanding is motivated by
`the fact that thermodiffusion in water might lead to powerful
`all-optical screening methods for biomo!ecules and colloids.
`Equally well, thermodiffu~ion handles and moves molecules all(cid:173)
`optically and therefore can complement well established methods:
`for example, electrophoresis or optical tweezers. For the latter,
`forces of optical tweezers scale with particle volume and limit this
`method to particlr..s of only >500 nm. Electrophoresis does not
`suffer from force limitations but is difficult to miniaturize because
`of electrochemical reactions at the electrodes,
`On the other hand. thermodiffusion allows the microscale
`manipulation of smali particles and molecules. For e.xamplc,
`1,000-bp DNA can be p~>HcrMd ;srbitrmf!y in bulk water (Fig, 1}.
`The temperature pattC'ru "DNA{ hea.tcd by:4 K, •va$ \>i:dtten into
`a water 'mm w[th an infntn::d lm>i?.t <ll~Mll.hlg micnmetip!~. Til~'
`concentration of 1,000-bp DNA was imaged by using a fluores(cid:173)
`cent DNA tag. In an overaH cooled chamber at 3°C, DNA
`accumulates toward the heated letters "DNA" (negative Soret
`effect), whereas at room temperature DNA is thermophobic
`(positive Soret effect) as seen by !he dark letters.
`In the past, the apparent complexity of therrnodiffus1on pre(cid:173)
`vented a full theoretical description. As seen for DNA in Fig. 1,
`molecules characteristically deplete from regions with an increased
`temperature, but they can also show the inverted effect and
`accumulate (2, 3). Moreover, the size scaling of thermodiffusion
`recorded bv thermal f1e.ld flow fractionation showed fractional
`power laws ~ith a variety of exponents that are hard to interpret { 4,
`5). The latter effect might be resolved by revealing nonlinear
`ibermovhoretic drift for the strong thermal gradients used in
`thermai field flow fractionation (our unpublished observations).
`A variety of methods were used to measure thermodiffusion,
`mostly in the nonaqueous regime, ranging from beam deflection
`
`- -20"C
`
`iOOpm
`
`fig. 1. Tnermodiffu>ion m~nipulata; the DNA concentration by small temp~r­
`atured!fferenceswrthinthe bulk$Oil,ticm.A thir.waterfilm i> heat~d by 2 K ;;long
`the letters "DNA" with an infrared la>er. F<:.>r a moled chamber at .3'C, tluo;,,.
`cently tagged DNA accumulntes at the w~•m letters. However, at room temper·
`ature, DNA moves Into thr.; cold, shewing reduced fluorescence. The ch~mber is
`60 I'-m thin, containing SO nM DNA i11 1 mM Tris buffer. ~ve•y 50th bi!S2 pair i>
`labeied with TOTO· 1 {for det~ils, see >upporling ir.fcrmation).
`
`(2, 3, 6), holographic scattering (7-9), electrical heating (10), to
`thermal lensing (11 ). Recently we have developed a fluorescence
`microl'luidic imaging technique (12, 13) that aHows tbe mea(cid:173)
`surement of thermodiffusion over a wide molecule size range
`without artifacts induced by thermal convection. Highly diluted
`suspensions can be measured; therefore, par!ide-particle inter(cid:173)
`actions do not have an influence. We only apply moderate
`temperature gradients. In !.he following study, we used this
`method to confirm a straightforward theoretical explanation of
`thermodiffusion.
`
`Theoretical Approach
`For diluted c.oncentrations, it is generally assumed (14) that the
`thermodiffusive drift velocity v depends linearly on the tempe.r·
`ature gradient VT with a proportionality constant which equals
`the thermodiffusion coefficient D-r: v = -DTVT. In steady state,
`thermodiffusion is balanced by ordinary diffusion, Constant
`diffusion and thermodiffusion coefficients both lead to an
`exponential depletion law (15) c(cu "' exp[-··(DT/D)(T - ?'o)J,
`with the concentration c dependmg on the temperature dtffer·
`ence T - To only. The concentration c is normalized by the
`boundary condition of the concentration co with temperature To,
`The Soret coefficient is defined as ratio ST = D·r/D, which
`determines the magnitude of thermodiffusion in the steady state.
`Although !he above exponential distribution could motivate an
`approach based on Boltzmann equilibrium statistics, it is com(cid:173)
`monly argued that thermodiffusion without exception is a local
`nonequilibrium effect that re.quires fluid dynamics, force fields,
`or particle-solvent potentials (16-20). However, in a previou>
`paper (15), we demonstrated that for moderate temperature
`
`Author contribution!: D. B. d~•igned r~seor<h; S.D. •od D.B. perform<!d rese•rrh; S.D. ond
`DJ3. anajy-ted data; ~nd 5-D. and D.B. wrot1=1 th~ pi!pt~:r.
`Th~ author~ di:!d:!l!re no umiHct of lntere:it.
`'To whom mrre5poc.de~« should!>~ ad~ressed. E-maH: diater.br•un~lmu.rla.
`C 2GDG by lhe NatfonaJ Acedemy of Sdene~~ of th~ USA
`
`w\o.<w.pna~.org/cgl/doi/1 0.1 073/pna;,06038731 OJ
`
`Page 1700
`
`TEVA EXHIBIT 1007
`TEVA PHARMACEUTICALS USA, INC. V. MONOSOL RX, LLC
`
`

`
`a
`
`Fig. 2. Sa it dependence. io) Thermodiffu;ion in water Is domlnat~a by kmk
`>hieldlng (Left) and water hydration (Right}. {i;) Soret weffici<;r;t Sr versus
`De bye ;.,,.,gth for carboxyl-'11odih~d poiy>tymne be~ds ot diaml.!\<:>r 1, 1, 0.5.
`and 0.21-'m. Une~r plot (Udt) ar;d logarithmic plot (,qigl>t). The Smet coeffl·
`dents are de;cribed by Eq. :<with an effective surface d1Mge of ,-•ff = 4,50(1
`ef1,m' ~nowr; from ele,·uophor<e~h. The intercept .Sr(~n~ = 0) ;, fitted with a
`hydration <entmpy p<>r part!de ;uoia;:~ of Sh;o = -1,400 Ji(moH<·;.<m<).
`
`ptay a role and we can introduce an effective surface charge
`density <Torr= Q.rr/A per molecule arell A. From the tempe.r(cid:173)
`ature detiva!ion according to Eq. 1. the ionic contribution to
`the Sor et coefficient is s¥""'") = (A {:ltr~rrA::m)/( 4esok T2
`). A
`similar relation was derived for charged micelles recently (22),
`although without considering the temperature dependence of
`lhe dielectric coeffic:e.nt e. Next, ;he. contribution to the Soret
`coefficient from the hydration entropy of water can be directly
`inferred from the particle-area-specific hydration entropy
`Soyrl = Soya/A, namely s~~yd) = -A,~~yd(T)ikT. Finally, lhe
`contribution from the Brownian motion is derived as ST = 1/T
`by inserting the kinetic energy of the particle G "' kT into Eq.
`1. However, this contribution is very small (ST = 0.0034/K) and
`can be neglected for the molecules under consideration. The
`wmribution~ from ionic shielding and hydration entropy add
`up to
`
`[2J
`
`{3fT~rr
`\
`-,l'oyd + -4
`-T--: X Am;l,
`seo
`/
`The Soret coefficient Sr scales linearly with particle surface A
`and Debye length Arm. We tested Eq. 2 by meusuring ST versus
`salt concentration, temperature, and molecul;; size. In all cases,
`thennodiffm;ion is quantitatively predicted without any fre-e
`parameters. We used fluorescence single-particle tracking to
`follow carboxyl-modified polystyrene beads (catalog no, F-8888,
`Molecular Probes, Eugen<J, OR) wittl diameters of 1.1 and 0.5 at
`25 ;;M dialyzed into 0.5 mM Tri~+J Cl at pH 7.6. Thermodiffusion
`of particle~; ::s0.2 ~-tm is measured by the fluorescenc.e decrease
`that reflects the bulk depletion of the particles (12). The
`chamber thickness of 20 !Lm d;;mped the thermal convection to
`negligible speed!i (15). The experimental design also excludes
`thermal lensing and optical trapping (15). Debye lengths Am:
`were titrated with KCI (see the supporting information, which is
`pubiished on the PNAS web site)o
`
`A ('
`Sr = -k--
`-;
`,
`
`1-
`
`Salt Di!pllndsnte. Fig. 2b show5 the Soret coefficients of polysty·
`r<Jne beads with different !lizes versus ,\Dfl· The So ret coefficients
`
`gradients the thermal nuctuations of :he part.lde are the ba5is for
`a loco:) equ11ibritsm. This al!ows the de:;criptlon of the thermod(cid:173)
`iffusive steadv state by a succession of local Boltzmann laws,
`yielding cleo "= exp[ -·(G(:r) - G(T0))ikT], with G being the
`Gibbs-free enthalpy of the single particle-solvent system. Such
`an approach is valid only if the aemperature gradient VT is below
`a threshold 'YT < (aSr)-l, which is given by the particle
`fluctuations ·,vith the hydrodynamic radius a and Soret coeffi(cid:173)
`cient ST, 1l5 shown recently (1.'5). In the pre:;ent study, temper(cid:173)
`ature gradients below this limit were used so that thermodiffu(cid:173)
`sion is measured at local thermodynamlc equllibrium conditions.
`Local thermodynamic equilihrium allows the derivation of a
`thermodynamic foundation of the Soret codficient. The local
`Boltzmarm distribution relates small concentmtion changes &
`with smBil Gibbs-free energy differences: &/c ""' -oG/kT. We
`equate ihiR relation with a locally linearized thermodiffusion
`steady state given by &/c "' -SroT and thus find the Soret
`coefficient by the temperature derivative of G:
`
`[1]
`
`Whereas the above relation is stJfficient for the following
`derivation, it can be generalized by locally applying the thermo(cid:173)
`dynamic rehiion dG = -SdT + v'dp + p.dN. For singi<J particles
`at a constant pressun:. we find that the Soret coefficient equals
`the negative entropy of the partide-solvmlt system S 1:ccording
`to Sr = -SikT, Thi~ relation is not surprising given that the
`entropy is by ddinition related with the temperature derivative
`of the free enthalpy.
`The above general energe1ic treatment is inh<Jrent in previ(cid:173)
`ously described approaches baoed on local equilibrium (14, 21,
`22), including the successful interpretatlon of thermoelectric
`voltages of diluted electrolytes (23, 24), which are described by
`energies of !nmsfer. Re{:ently, the nonequilJbrium approach by
`Ruckenstein (25) was applied to colloids (26) with the charac·
`teristic length I as:;igned to the De bye length ADH· If in~tead one
`would assign the charaet<Jri;;tic length according to l = 2ai3 with
`the particle radi!JS a, the Rucke.nstein approach would actually
`confirm the above local equilibrium relation (1) for the Sorel
`coefficient Measurements em SDS micelle~ (26) appeared to
`confirm this nonequilibrium approach, but for the chosen par·
`tides the competing parameter choices l ='J.a./3 and C = Arm
`yielded comp::~rabfe value~. Thus, the experiments could not
`distinguish between the competing theories.
`We will use the above local equilibrium relation;; to derive
`the Sore: coefficient for particles larger th'm the Debye length
`in aqueous solutions and put the results to rigorous experi·
`mental tests, Two contributions dominate the panicle entropy
`Sin water (Fig. 74): the entropy ofionic ~bidding (Fig. 2a Lejt)
`and the temperature-sensitive entropy of water hydration {Fig,
`2a Right). The contribution from the entropy of ionic shielding
`is calculated with l he tempenltur<J derivative of the Gibbs" free
`enthalpy (26, 27) Giooio "' Q~rrArm/[2A eeo] with the effective
`charge Qerr and particle surface A. Alternatively, this enthalpy
`can be interpreted as an electrical field energy G;.,,i<• "'
`Q~u![2C] in the ionic !lhielding capacitor C. We neglect the
`particle-particle interactions because the fluorescence ap(cid:173)
`proach allows the measurement of highly diluted systems. To
`obtain the Soret coefficient, temperature derivative~ consider
`the Debye length Arm(!} = v'i(T)s-;;k;i?(ze2cs) and the
`dielectric cor:stant t:(T), Bo:h remperalure derivatives give rise
`to a factor f3 = 1 -
`('T/e)ae/BT, The effective. charge Oorr is
`largely temperature-insensitive, which was confirmed by eke(cid:173)
`trophores.is independently (28). Such a dependence would be
`unexpect<Jd because the strongly adsorbed ions dominate the
`value of the effective charge. Experimentally, we deal with
`colloids exhibiting flat surfiices, i.e., the particle radius is
`larger than Atm- In this case, charge renormalization does not
`
`Duhr and Braun
`
`Page 1701
`
`TEVA EXHIBIT 1007
`TEVA PHARMACEUTICALS USA, INC. V. MONOSOL RX, LLC
`
`

`
`0
`
`10
`6
`Debye Le•1gth ;~0 , [nm]
`
`fig, 3. Temperature d<Jpend;,rKe. !a) Th<> t<>r>•perah:ra dependencl.' is don•·
`!n,;tr;d by the linear {hange In tM hydmtion entropy Sevd· It shifts the ~alt·
`dependentthermodiffu;ion Sr(AoH) to lowe,value>. The partlde si>'e i:; L 1 !!m.
`(b) Th~ Smet w.zffident $, inm.>asl.'> linr;arlywith th<>temperature ;;s expect<Jd
`for a hydratkm entrop~; 5nf0(Y). i; depend; or: the molewl;; >p>;cie;, not it; ;lze.
`a:; se€!r; from the rescaled 5cret coefficients for DNA with different length>.
`
`scale linearly with a small intercept at ADH = 0 and confirm the
`ADwdependence of Eq. 2. For smaller·diameter beads, the Soret
`coefficients scale with the particle surface area A (Fig. 2), as
`expected from Eq. 2. To check whether Eq. 2 also quantitatively
`explains the measured Soret coefiklcnts, we inferred the effec~
`tive charge of the beads by electroptlmesi:! (see supporting
`materials). By using 40-nm beads with identical carbOJ..')'l surface
`modifications at ADH = 9.6 nm, we nuorescently observed
`free-flow electrophoresis and l!orrected for electroosmosis, find·
`ing an effective ;;urface. charge density of o-0rr = 4,500 ± 2,000
`e/p..rn2• Thi.:; va!ue is virtually independent from the used salt
`concentra1kms (28). With rhis inferred effective charge, Eq, 2 fits
`the Soret coefficient for various bead sizes and salt concentra(cid:173)
`tions weil (Fig. 2b, solid lines).
`The intercept ST(ADH = 0), where ionic contributions are zero,
`also scales with particle surface and is described by a hydration
`entropy per particle surface of Shyd = -1,400 J/{mol·K·~<-m2). The
`value matches the literature values for similar surfaces reason(cid:173)
`ably well (29-31). For example, dansyl-alanine, a molecule wlth
`surface groups comparable with polystyrene beads, was mea(cid:173)
`sured to have a hydration entropy (29) of -0.13 J/(moi·K) at a
`comparable temper<lture. Linear scaling with its surface area by
`using a raditw of a "" 2 nm results in a value of Snyd = -2,500
`J/(mol·K·p..m2), in qualitative. agreement with our result The
`hydration entropy is a highly informative molecule parameter
`that is notoriously difficult to measure, yielding an interesting
`application for thermodiffusion.
`
`'hlmperaturl! Oep~ndencl!. Hydration entmpie:; Snyd in water are
`known w increa~e linearly witb decreasing temperatures (29-
`31). Because the slope of the ionic contribution of Srvenus A.m1
`ls with high-precision temperature insensitive for water [f3(T)I
`(eT2) s const), only the intercept is expected to decrease as the
`overall temverature of the chamber is reduced, Thi~ is indeed the
`case, as §een from the temperature dependence of beads with
`diameters of L1 ,um (T = 6-'29"C) (Fig. 3a). We infer from the
`intercept S·r(ADfl "' 0) that the hydration entropy changes sign at
`=20°C. As seen for DNA in Fig. 1, hydration entropy can
`dominate the.rmodiffusion at low tempera:ures and move mol(cid:173)
`ecules toward the heat (O.r < 0).
`The properties of hydration entropy lead to a linear increase
`of S.r over temperatures at a fixed aalt concentration as measured
`for 1.1-f<.m beads and DNA (Fig. 3b). We normalized Sr by
`dividing by .h(30°C) to compensate for molecule ~urface area.
`The slope;; of ST over temperature differ between beads and
`DNA, Howeve.r the slope. does not differ be!ween DNA of
`different size (50 bp versus 10,000 bp). Based on Eq. 2. this is to
`
`be expected because !he temperature dependen~"' of the hydra(cid:173)
`tion entropy depends only on the type of surface of the molecule,
`not its size. We measured the diffu;;ion coefficients of the DNA
`species at the respective tempilr<:ture independently. Witbin
`experimental error, changes in the diffusion codficient D match
`with the change of the ·.vater viscosity without the need to assume
`conformationul changes of DNA over the temperat!lre range.
`Please note that the change of the sign of the DNA Soret
`coefficient is situated near the point of maximal water density
`only by chance. There, the two entropic contributions balance,
`For polystyrene beads at Arm ~ 2 nm for example, the sign
`change is observed at 15°C (Fig. 3a ). An increased Soret
`coefficient over tempemture was reported for aqueous solutions
`before (3), however with a distinct nonlinearity that we attribute
`to remnant particle-particle interactions.
`
`Siie Oeptnds;nc!! of the !lead$, The Soret coefficient was measured
`for carboxyl-modified polystyrene beads in diameters ranging
`from 20 nm to 2 /Lm, Bead:; witb di<lme!ers of 0.2, 0.1, O,G4, and
`0.02 iJ.rH were diluted to concentrations of 10 pM, 15 pM, 250
`pM, and 2 nM, and their bulk fluores(:ence was imaged over time
`to derive D7 and D (12, 15) from the depletion and sub:;equent
`back-diffusion. Larger beads with diameters of 1.9, 1.1, and 0.5
`~·m were diluted w concer!!rlltions of 3.3 aM, 25 aM, and 02 pM
`and measured with single-particle tracking. The solmiom were
`buffered in 1 mM Tris (pH 7.6) with Aml = 9,6 nm. ln all cases,
`interaction§ between particles can be excluded. Care wa:s taken
`to keep the temperature gradient in the local equiHbrium regime.
`We find that the Soret coefficient §Cales with particle surface
`over four order~ of magnitude {Fig. ·1a). Tbe data are described
`well with Eq. 2 with an effective surface charge density of rrorr =
`4,500 e!~<-m' and neglected hydration entropy contribution. The
`5-fold too-low prediction for the smallest particle {20 nm in
`diameter) can be explained by charge renormalization because
`its radius is smaller than itDB·
`The diffusion coefficient D for spheres is given by the Einstein
`relation and scales inversely with mdiu:; D a< 1/a. In~erting Eq. 2
`into ST "" Dy/D, the thermodiffusion coefficient DT is expected
`to scale with particle radius a, This is experimentally confirmed
`over two orders of magnitude (Fig. 4b ). These findings contmdict
`several theoreitcal studies claiming that DT should he indepen(cid:173)
`dent of particle size (16-20, 26), based on ambiguous experi(cid:173)
`mental resul!s from thermal field flow frac:icmation (4) thal
`were probably biased by nonlinear thermodiffusion in large
`thermal gradients (15).
`
`Si~i! Oeps;m:lsm::t of ONA. Whereas polystyrene beads share a very
`narrow size distribution as a common feature with DNA mole(cid:173)
`cules, bends are a much tess complicated model system. Beads
`are rigid spheres that interact with the solvent only at its surface.
`ln addition, the charges reside on the surface, where the
`screening takes place. Thus, the finding that thermodiffusion of
`flexible and homogeneously charged DNA is described equally
`well with Eq. 2 is not readi!y expected and quite interesting (Fig,
`4 c and d).
`We measured DNA with sizes of 50-48,502 bp in 1 mM Tds
`buffer {ii.DH = 9.6 nm) Ht !ow molecule (:oncer:trations be1ween
`1 j.tM (50 bp) and 1 nM (48,502 bp). Only every 50th base pair
`was stained with the TOT0-1 fluorescent dve. The diffusion
`coefficient was measured by back-diffusion after the laser was
`turned off and depends on the length L of the DNA in a
`nontrivilll way. The data are well fitted with a hydrodynamic
`radius scaling a :x L 0·7·5• This scaling represents an effective
`average over two DNA length regimes. For DNA molecules
`longer than =1,000 bp, a scaling of 0.6 is found (32), whereas
`shorter DNA scales with an exponent of =1 (see the supporting
`information),
`We can tfescribe the measured Soret coefficient over three
`
`Duhr am:l Bfilur.
`
`Page 1702
`
`TEVA EXHIBIT 1007
`TEVA PHARMACEUTICALS USA, INC. V. MONOSOL RX, LLC
`
`

`
`::r
`
`"! :ll
`
`71
`.---/ I
`
`.. ~~~~'>••~-:;H'~{.~'~r
`c~
`0.1
`1
`oo~
`r:iam~!ac lurr.]
`
`w'
`1c'
`10'
`DNA Lenglo [bpJ
`
`Si?.e droprondenc<'. (a) for poly>tyrene bead,, the Soret wefflciant
`fig. 4.
`"'al~s witr, the particle !iUt'face ave ;four ord;:r; ot magnitude. Mea;uri!ments
`are described by Eq. 1 with an effective surface charg;; demit;.· of .,..rr = 4,500
`ei;m/ {2) and negligible hydration ;;ntropy. Th~ deviation forth<: bei!d wi•h
`3 dianwter of 20 r.m car; be omd<mtoodfrom ~n increa;ed effective charge due
`to the onset of chargE normali<:~tior. for a :;;,<.01~· (b) Acmrdir>giy, t!w ther(cid:173)
`modiffu;ion r.cefflclent Dr scale> linearly with bo:ad diarnet'<'r. (c) The Scret
`((:!!ffider.t (Jf DNA >cal<os ar.wrding to Sr "''/i.., with the iength L of the DNA
`based on EQ. 2 with <on e!tecrive charge per base pa>ir oHU2 e. (d) Thermod·
`iffu>icn coefficient Dr deuea>e~ over DNA length wilh 0,·" L -<'·'", <:dlJ><OrJ by
`the $t~aling of diffusion rceffident D •> L -v>.
`
`order;; of magnitude of DNA lengths with Eq. 2 if we assume that
`effective charge of the DNA is shielded at the surface of a sphere
`with 1he hydrodynamic radius a. Because of the low salt con(cid:173)
`centration (Amr = 9,6 nm), such globular shielding is reasonable.
`Not only is the experimentally observed scaling of tbe Sore1
`coefficient with the squa~e root of itg length correcdy predicted
`based on Eq. 2 (Sr :X o:rda2 o: U!V 5 IX L05), but the Soret
`coefficient also is fully described in a quantitative manner (Fig.
`4c, solid line), with an effective charge of 0.1:2 e per base,
`matching we!( with li!erature values (33) ranging from 0.05 e/bp
`to 0.3 etllp.
`As shown in Pig. 1kl, the thermodiffusion coefficient for DNA
`drops with DNA length ncconling to DT = DS;: c: Q~rda 3 oc
`L2ff-2·25 oc L -0.25. Thu~, shorter DNA a:::tuaHy drifts faster in a
`temperature gradient than longer DNA. It is important to point
`out that this finding is in no way contradictory to experimental
`findings of a constant DT over polymer length in nonaqueous
`settings (8), According to Eq, 1, the thermodyno1mic relevant
`parameter i:; !be Soret coefficient, which is determined by the
`solvation energetics, The argument (J 9) that polymers have to
`decouple imo monomers to show a constant DT merely becomes
`the special case where the solvation energetics determine both
`STand D with equal but inverted size scaling. In accordance with
`our local energetic equilibrium argument, Sr and not DT dom(cid:173)
`inates thermodiffusion also for nonaqueous polymers near a
`glass transition (8). Here, Sr is constant, where.as D-r and D scale
`according to an increased friction. However, for a system of
`DNA in solution, for which long-ranging shielding couples lhe
`monomer>, a constant DT over polymer length :::annN be as ..
`sumed a priori (Fig. 4d).
`
`~Effective C!mrge, The effective charge Q0 ;r is a highly relevant
`parameter for colloid ~cien:::e, biology, and biotechnology, So far
`it only could be inferred from electrophoresis, restricted to
`
`0. i
`O.QOj O.o1
`fla!ticla Suri<~ce !i->m2J
`
`10
`
`EH~'tive d1arge from thermod;Hmion. Effett!ve rharge i; interred
`Ftg. !>,
`from thermodiffusion u;ing Eq. 3. Poly;;tyre11e beads (:;tQ .. ;;t,OOO nm) (a) and
`[l~U\ (S.0-50.000 bp) (b) W<.'re mea>ur~d ever a l;~rge ~i&e r<mge, whkh i:;
`impn~>ible wi·th elertrophore:;i,, A> e~fJf!(.ted, the effectl11e r.h<orge nf the
`bead~ s{aie> with partids >Urface <ond lineariy with the length of DI\JA.
`
`part1clcs smaller than the Debye !engt!J (a :s 3..\nH) (34),
`Unfortunately, many colloids are outside this regime, As shm•m
`before, a similar size restriction does not hold for thermodli'fu"
`skm. In many case5, the hydration entropy Snyd comribmes <15%
`(Fig. 2) and can be neglected at moderate salt levels. Thus, we
`c;m invert Eq, 2 to obtain the el'feclive charge Q"rr for spherical
`molecules from
`
`[3]
`
`The effective charge derived from thermodiffusion measure"
`ments of polystyrene beads and DNA is plotted in Fig. 5 over
`;everal orders of magnitude in size, The effective :::hBrge of beads
`~:::ale;; linearly with particle ~urfacc, with a slope confirming the
`effective surface charge density of if err= 4,500 el~&m 2, which was
`inferred from electrophoresis only for small particles, Average
`deviations from linear scaling are <8o/! .. (Fig. 5a). The effective
`charge inferred from thermodiffusion mea:mrements of DNA
`using Eq. 3 scales linearly with DNA length with an effective
`charge of 0.12 e/bp. The length scaling is eonfirmed over four
`orders of ma.gnitude wiih an average error of 12% (Fig . .'ib).
`Thus, thermodiffusion can be used to infer the effective charge
`with low error~ for a wide range of particle sizes" This is even
`more interesting for biomolecule characteri:mtion because mea"
`surements of thermodiffusion can be performed all-optically in
`picoliter volumes,
`
`Com:lsa!>ion
`We describe. thermodiffusion, the molecule drift along t;;mper(cid:173)
`ature gradients, in liquids ·,vith a general, microscopic lheory.
`Applied to aqueous solutions, this theory predicts thermodiffu·
`:;ion of DNA and poly:;tyrene beads with an average accuracy of
`20%. We experimenta!ly validate major parameter dependencies
`of the theory: linearity against screening length f..m1 and mole"
`cule hydrodynamic area A, quadratic dependence on effective
`charge, and linear1ty against temperature, Measurements of
`thermodiffusion can be miniaturized to the micrometer :;:::ale
`with the all-optical fluorescence lechnique ;md permit micro·
`acopic temperature diffe.rences to manipulate molecules based
`on their surface properties (Fig. 1 ), The theoretical description
`allows the extraction of §olvation entropy and the effective
`charge of molecules and particles over a wide :>ize range,
`
`Mcsteriai!i and Method!>
`!nfrars:d Temps:ratur!! Col'ltm!. The tempeBture gradients used to
`induce thermodiffusive motions were created by aqueous ab(cid:173)
`sorption of an infrared laser (Furukaw<: Electric, Tokyo, Japan)
`at a wavelength of 1,480 nm and 25 mW of power. Water strongly
`
`Page 1703
`
`TEVA EXHIBIT 1007
`TEVA PHARMACEUTICALS USA, INC. V. MONOSOL RX, LLC
`
`

`
`ubsorbs at this wavelength with an attenuation length of K "' 320
`p.m. The laser beam was moderately focused with a lem of 8-mm
`focal {li:;!ance. '(vpk:aHy, lh~ Hmlp(mtl<H\~ i.n rhtl S<ibtion 'IV!~~;
`raised by 1-2 Kin the bcum cenrerwllh a l/~~~di;l.l11»t<~rot25 t<.m,
`measured with th,~ ttmp~mlur{:·d(:p~~niknl f!m.'>!t~:>.CelJcc s.i~ri~d
`of the dye 2' ,7' obJS(Carl:m:X)'t::lh)'l}o5(fl)·~~"rbo~yfh#JI"(:S(.~CitJ (12).
`Thin chamber heights of l0-20 t<-m and moderate focusing
`removed posRible artifacts from optical trapping, thermal lens·
`ing, and thermal convection (12-). For tempera!ure-dependent
`measurements, both the objective and the microfluidic chip were
`tempered with a thermal bath. Imaging was provided from an
`A:;;ioTech Vario fluorescence microscope (Zeiss, Oberkochen,
`Germany), illuminated with a high·power light-emitting diode
`(Luxeon, Calgary, Canada), and recorded with the CCD camera
`SensiCam QE (PCO, Kelheim, Germany).
`
`Mu~ew~e~. Highly monodisperse and protein-free DNA of 50,
`100, 1,000, 4,000, 10,(}00, and 48,502 bp (Fast Ruler fragments
`and >.-DNA; Fermentas, St. Leon-Rot, Germany) were diluted
`to 50 i;.M base pair concentration, i.e., the molecule concentra(cid:173)
`Hon was between 1 '";Jvi (50 bp) and l nM ( 48,502 bp ). DNA was
`fluore~cently labeled by the intercalating TOT0-1 fluore~cent
`dye (Mole.cu!ur Probes) with a low dye/base pair ra;io of 1:50.
`Carboxyl-modified poiysty rene beads with diameters of 2, 1, 0.5,
`0,2, OJ, 0.04, and 0.02 p.m (cat,;!og nos. F-8888, F-8823, F-8827,
`F-8888, F-8795, F-8823, and F-8827; Molecuiar Probes) were
`dialyzed (Elutll Tube mini; Ferment;>!i) in dlstHled water and
`diluted in 1 mM Tris (pH 7.6) to concentrations between 3.3 aM
`(2- ILm) and 2 nM (0.02 J.l.m).
`
`Co!lcrc>r~tr<~thm !m<~gl!ig Ovr<>r Thni!, Ei1her the method of concen(cid:173)
`tration imaging (12) or single-particle tracking was used tD
`
`measure thermod!ffu~ion at low concentralions, narne,Jy <0.03
`g!Hter for DNA and 10-5 g_niter for beads. At higher concen(cid:173)
`trations, we found profound changes of thermodiffusion coef(cid:173)
`ficients, DNA and polystyrene beads of <0.5 !L!ll in diameter
`were imuged over time (12} by brlght·field fluorescence with a
`X40 oil-immersion objective. Concentrations inferred after cor(cid:173)
`recting for bleaching, inhomogeneous mumination, and temper(cid:173)
`ature-dependent fluore:;cence (12) were fitted with a finite
`element theory, The model captures all de!ail$ of both ther(cid:173)
`modiffusive depletion and back-diffusion to measure D,. and D
`independently (aee 5upporting informa;t!on). Measurement>
`were performed in mkrof!uidic ch!ps 10 p.m in height with
`polydimethylsiloxane on both. sides.
`
`Slngi&-!'>;srtkle Tracking. Polystyrene particles of >0.5 p,m in
`diameter were measured by single-particle tracking due to the
`slow equilibration time and risk that steady-state depletion is
`disturbed by thermal convection. The thermodiffm;ive drif! was
`imaged with a x32 air objective at 4 H:r. at an initial stage of
`depletion in a 20-~m-thick chamber, Averaging over the z
`position of the particles removed effects from thermal convec(cid:173)
`tion. The drift veiocity versus temperature gradient of 400 tracks
`were linearly fitted by v = -D,.VT to infer DT. The diffusion
`coefficients D of the partldea were evaluated based on their
`squared displacement, matching wirhin 10% the. Einstein
`relatiollship.
`
`We thlmk Klam Stierstadt, Jan Dhorn, and WenJ<:r Kohl<~r for digcus(cid:173)
`sions and Julia Morfill and Ve,ronil:a Egger for comments on the
`manuscript. Our Emmy-Noether Group is funded by the Deutschr.
`Forsdmng:;gemeinschaft and hosted by Hermann Gaub.
`
`L L~dwig C (1856} Si!mngber Bayer Alwd W1Ss !!1en Math-l'ialwwiss K/20:539.
`2, Giglk) M, Vendramil!i A (1977) Ploys Rev Lei/ 38:26-30.
`l Jacopi11i S, l'i&ua R (20lB} Europhys Leu 63:247--253.
`• :, Jean SJ, Sd;impf ME, Nyborg A (1997) AMI Chem 69:3442--3450,
`5. Shimdu PM, Lw G, Giddings JC (1995) Anal Chern 67:27(!5-27!3.
`6. Pi:u~a R, Guarino A (2002) l'hys RE'' Lei/ 88:208302,
`7. de Gan~ BJ, Kita R, Muller B, Wieg<md S (2003) J CMm Ploys ll8:8073--SO&l.
`B. Rauch J, Kilhl~r W (20(!2) Phys Re:• Lea 86:185901.
`9. \Vif.gand S1 Kt.'lhh:~r Vl (2002) in Th.~rmal Nunequilibriu.;!'s Phenarm.mtJ ln Fluid.
`Mixlures (Springer, Il~rlin), p 189.
`10. Putnam SA, C:ilim DG {2005) Langmuir 21:5317-5323.
`11. Ru"cor.i R, !saL, Piaw! R (1004) J Opl Soc: Am B 21:605--616,
`12. Dul:r S, AJduini S, B.-sun D (20{)4) Ew P!:ys J E 15:2.77--286.
`l3 Braun D, Lit,.:hab~; A {2002) Phy.r Rtl' Leu 89:188103.
`l4, de Groat SR, Ma<ur J' {l %9) Nonequiiibrium Thermodynamics (North-
`Hol!and, Amstm:lam),
`lS. Duhr S, Bra"n D {2006} Phys Re" Le!i 96:1()3301.
`l6. Emery AH, Jr, Driclamer HG (1955) J Chem l'hys 23:2252--2257.
`i7. H?.m J (l96G) J Appi Phy.r 31:1853-1858.
`18. Mon.m:>Y KI {1999) J E<p Th•or Phys 88;944-946,
`
`19. Schimpf ME, ikmemw SN {2.000) J l'hys Chern 8 104;9935-9942.
`2(). Vail A, Krdrhrlv A, Enge W, Kramer L, Kl\hlel W (2005) Phys fley L<ll
`92:214501 .
`2l. Dhom JKO (201Yl) J CMm l'llys l2():l632-:l64:l.
`2·2. F~ycll~ S, Bkk,~l T, l

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