Finite-size scaling and particle size cutoff effects in phase separating polydisperse fluid

合集下载

英语翻译

英语翻译

Moisture and Strain Rate Effects on Concrete Strengthby C. Allen Ross, David M. Jerome, Joseph W. Tedesco, and Mary L. HughesThe effects of strain rate and the coupled effects of moisturecontent and strain rate on concrete strength are experimentallyinvestigated. Dry concrete exhibits clearly defined strain rate tran-sition regions for both tension and compression. Significantstrength increases in dry concrete are observed only at strain ratesabove the transition regions. Wet concrete, however, experiencesappreciable increases in strength at strain rates both below andabove the dry concrete transition region. An analytical expres-sion has been derived that accura tely predicts the tensile fracturestrength of dry concrete. This expression can be modified toaccount for moisture content in concrete.Keywords: compressive strength; concrete s; moisture content (of aggre-gate,hardened concrete, etc.); strains; tensile strength.INTRODUCTIONResults from experimental studies documenting theeffects of strain rate on dry concrete strength have beenreported in the literature as early as the mid-1900s.1-3 Sincethat time, a considerable amount of data relevant to the strainrate dependence of concrete strength has been published.Specific details of some of the methods used for data collec-tion are presented in References 4 through 14. Recent workconducted by Ross et al.15 demonstrates strain rate sensi-tivity in both compression and tension for various dryconcrete mixes. Compressive strength data for dry concreteis presented in Fig. 1. This figure illustrates that no appre-ciable strength increases occur below a strain rate range ofapproximately 60 to 80/sec. Beyond this transition region,however, significant increases in strength are exhibited.Similar data for concrete tensile strength is presented in Fig. 2.The strain rate transition region for appreciable increases indry concrete tensile strength is approximately 1 to 10/sec.The presence of excess moisture in concrete appears tohave a significant effect on both the tensile and compressivestrengths under high strain rate loading. Severalinvestigators15-17 have recently reported strength increasesin wet concrete greater than those exhibited by dry concretefor strain rates above the tran sition region. Moreover, Rossiand Boulay16 have observed substantial strength increasesfor wet concrete at strain rates below the transition region,where dry concrete has been demonstrated to be relatively insensitive. Based on their observations, Rossi and Boulay16have concluded that significant strain rate effects exist in wetconcrete specimens only. Ho wever, based on concretecompressive strength data presented by Ross et al.15 andtensile strength data presented in this paper, Rossi andBoulay’s conclusion is valid only for strain rates below thetransition region. At strain ra tes above the transition region,both wet and dry concrete stre ngth data exhibit significantstrain rate sensitivity.Dynamic fracture stress, time of fracture, and particle sizeof brittle materials as a function of strain rate are given indetail by Grady.18-20 However, these theoretical relation-ships are only applicable above some high strain rate tran-sition region that is not defined. In this paper, the relationsof Grady18-20 are reviewed for application to concrete. Inaddition, a relationship is derived to predict dry concretetensile strength from very low to high strain rates.RESEARCH SIGNIFICANCEThe results of this research demonstrate clearly definedstrain rate transition regions for the compressive and tensileACI Materials Journal, V. 93, No. 3, May-June 1996.Received February 10, 1994, and revi ewedunder Institute publication policies.Copyright © 1996, American Concrete Institut e. All rights reserved, including the makingof copies unless permission is obtained fr om the copyright proprietors. Pertinentdiscussion will be published in the March-April 1997 ACI Materials Journal ifreceived by December 1, 1996.ACI member C. Allen Ross is a professor emeritus of aerospace engineering, mechan-ics, and engineering sciences, University of Florida, and a visiting professor at HQAir Force Civil Engineering Support Agen cy, Tyndall AFB, Florida. His researchinterests include the dynamic respon se of materials and structures.David M. Jerome is a senior research engineer in Wright Laboratory’s ArmamentDirectorate at Eglin AFB, Florida. He recently obtained his doctoral degree in engi-neering mechanics at the Univ ersity of Florida. His rese arch interests include thedynamic response of ma terials and structures.ACI member Joseph W. Tedesco is Gottlieb Professor of Civil Engineering, AuburnUniversity. His research interests include structural dynamics and finite element analysis.Mary L. Hughes is a research structural engineer in Wright Laboratory’s A ir BaseSurvivability Section at Tyndall AFB, Florida. She has received BS and MSdegrees in civil engineering.strengths of dry concrete. Below these transition regionsobserved strength increases are negligible; above thesetransition regions significant strength increases are exhib-ited. The moisture content of concrete has been shown tohave an important effect on strain rate sensitivity. Wetconcrete has displayed significant increases in strength atstrain rates below the dry conc rete transition region, and atstrain rates above the transitio n region, the strength increasesobserved in wet concrete are greater than those experiencedby dry concrete. This become s significant for submergedstructures when subjected to dynamic loadings that inducestrain rates above 1.0/sec. Finally, an analytical expression isderived from fracture mechanics theory that accuratelypredicts the tensile fracture strength of dry concrete over awide range of strain rates.Brittle fracture of concreteStudies on brittle fracture of condensed matter conductedby Grady18-20 are based on the assumption that local kineticenergy plus strain energy must exceed or equal the fracturesurface energy (all on a unit mass basis) for fracture to occur.Grady’s general assumption implies that at high st rain rates all of the particles of the same size must be fractured at thesame time. The basic premise is that peak stress σs is attainedin finite time ts at some strain rate and is given by(1)where B is bulk modulus, ε is strain, is strain rate, ρ isdensity, C0 is wave speed , ts is time to fracture, andσs is tensile fracture stress. At high strain rates, the fractureparticle diameter s is assumed to be limited by the wavespeed tos ≤ 2 C0ts(2)where the factor 2 comes from the assumption that a crackone-half length (s/2) = C0ts will be propagated in each directionfrom a fracture site.Details of the derivation of energy terms are presented inReferences 18 through 20 and are reviewed here. All energyterms are reckoned per unit mass andε·σsB ερC02ε·ts==ε·B ρ⁄ ()(3)(4)(5)where KIC is fracture toughness. Using the inequality KE + U ≥Γ(6)and Eq. (1) and (2), brittle fracture or spall strength σs, spallor fracture time ts, and fragment size s are found to beKinetic energy KE1120-------- - ρε·=2s2Strain energy density Uσs22 ρ C02--------------- =Fracture surface energy Γ 3 =KIC2ρC02s-------------- -(7)(8)(9)The strain energy density term of Eq. (6) has been shownto be approximately 15 times that of the kinetic energy term.Therefore, the kinetic energy term was neglected inobtaining Eq. (7) through (9).Eq.(7) through (9) are derived using the equality condition ofEq. (2) where the minimum particle size is limited by thewave speed. It has been shown experimentally15 that thecrack velocity of concrete is a function of strain rate. Thetheoretical limiting crack velocity is shown by Broek21 to be0.38 CL where CL is the longitudinal wave velocity. From Eq.(7) the slope of log10σs versuslog10εbecomes 1/3 . This slopeagrees reasonably well with the slope of regression curvesfitted to experimental data for concrete above a certain tran-sition strain rate.σs3 ρ C0KIC 2ε·() =13⁄ts1C0----- -3KICρCε·⁄()23⁄=s 2 =3KICρC0ε·⁄()23⁄However, Eq. (7) through (9) alone do not permit a contin-uous transition from low to high strain rates. Therefore, alow strain rate brittle fracture expression is introduced, usingcrack velocity as a function of strain rate, to facilitate thetransition from low to high strain rates.Low strain rate brittle fractureCrack velocity in concrete has been shown experimentallyto increase with strain rate15 (Fig. 3). Based on this data, therelation between crack velocity uc and strain rate ε.may bepresented as(10)where k is the intercept of the log10uc axis and m is the slopeof the linear plot of log10uc versus log10. Using Eq. (10),Eq. (2) may be rewritten as(11)tThe fracture toughness KIC that appears in Eq. (7) through(9) is assumed to be the static value and remains constantwith strain rate. Dynamic toughness KID, defined as thedynamic material resistance that is dependent on crackspeed, is given by Anderson22 as(12)where uc is crack velocity, ucl is limiting crack velocity, n isan experimentally determined value of approximately 1/ 3,and KIA is crack arrest toughness, defined in the limit as uc = 0,KID = KIAArrest toughness KIA appears to be approximately 25 to 50percent of KIC. The fracture toughness KIC of concrete isgiven by John and Shah23 as(13)where KIC is expressed in MPa-m1/ and fc′is the compressivestrength in MPa. The limiting velocity ucl has been shown tobe 0.38 CL by Broek21 and is assumed to approach Rayleighwave velocity CR as proposed by Anderson.22 Rayleigh velocity is approximately 0.6 CL for a Poisson’s ratio of ing Eq. (1), (10), (13), (3 ), (4), (5), and (6), anexpression for tensile fracture strength of concrete thatincorporates both crack velocity and fracture toughness asfunctions of strain rate is obtained. This fracture strength isgiven asKID KIA1 ucucl⁄ ()n–[] ⁄ ==ACI Materials Journal/May-June 1996(14)and is a function of strain rate, but most of the other param-eters are constants depending only on the compressive strength.Eq. (14) was evaluated using the following parameters:k = 100 (Reference 15)MPa (4060 and 8270 psi), were examined and the resultingstresses were normalized with respect to the stress at =10–5.7/sec, which yields the ratio of dynamic to static tensilestrength. The resulting curves are plotted versus log10 inFig. 2, along with experimental data. Very good correlationbetween Eq. (14) and the experimental concrete tensile datawas obtained. It is worth noting that very similar results wereobtained by Weerheijm25 using a much more elaborate and complicated derivation.It is important to point out th at a singularity occurs in thedenominator ofEq. (14) with the relationship of(15)–This singularity, which would limit the strain rate rangeover which Eq. (14) would be valid, can be overcome bylimiting the tensile strength to some peak value above thecritical strain rate given by(16)The advantage in using the ratio of dynamic to staticstrength, based on the low strain rate at which the static datawas obtained, is that the tensile strength at any strain rate isdetermined by simply multip lying the ordinate of thenormalized curve by the static tensile strength. This elimi-nates the necessity of having to know the arrest fracturetoughness of the concrete.As illustrated in Fig. 1, the compression data exhibits alow to high strain rate transitio n at a strain rate of approxi-mately 10 times that observed for the tensile data presentedin Fig. 2. This high strain rate compressive data was obtainedfrom experiments conducted on a 51-mm-diameter splitHopkinsonpressure bar (SHPB) using 51-mm-long specimensof the same diameter. (See ne xt section for a descriptionof the SHPB.) On failure these specimens experienced longi-tudinal cracking, thereby su ggesting transverse tensilefailure. Assuming the longitudinal cracking is the result ofstrain due to the Poisson effect, the transverse tensile strainεtt is computed byε·uclk⁄ ()1 m⁄=For a given Poisson’s ratio of 0.2 for concrete, both theaverage transverse tensile strain and strain rate are approxi-mately one-tenth of the compressive values. This results in acompressive transition strain rate that is 10 times the tensiletransition strain rate.Moisture and strain rate effectsIn an effort to experimentally determine the effects ofmoisture coupled with strain rate on concrete strength, aseries of wet/dry tests at low and high strain rates wereconducted. Tests were performed on completely wet, one-half dry, and dry concrete specimens.Concrete blocks of 0.3 x 0.3 x 0.3 m (12 x 12 x 12 in.) werecast, cured in air for 24 hr, placed in tap water for 7 to 10days, and then cored and cut to size underwater. The 51-mm-diameter, 51-mm-long (2.0 x 2.0-in.) specimens were storedin water for a total of 30 days.Previous tests15 have been completed to determine theoven time required for completely dry and one-half-dryspecimens. Oven drying at 230 F (110 C) with successivetimed weighings produced both the dry and one-half-dryspecimens. Wet specimens remained completely submergeduntil testing.Static tests were performed on all three types of specimensusing a closed-loop material test machine. Dynamic testswere accomplished using the Wright Laboratory SHPB atTyndall AFB, Florida. All tests specimens were 51 mm in diameter and 51-mm long. The split Hopkinson pressure baris a high strain rate device c onsisting of a gas gun, striker bar,incident bar, specimen, and transmitter bar, as shown inFig. 4(a). In general, the striker bar, propelled by the gasgun, impacts the incident bar and produces an elastic stresswave that propagates to the specimen, sandwiched betweenthe incident and transmitter bars. The elastic stress waveimpinges on the specimen wher e a portion of the stress waveis reflected back into the incident bar and a portion of thewave is transmitted into the transmitter bar. Strain gagesmounted on the incident and transmitter bar record the inci-dent and transmitted stress pulses. The integrated reflectedpulse is proportional to the strain in the specimen and thetransmitted pulse is proportional to the stress in the spec-imen. Using these pulses, a d ynamic stress-strain curve maybe formed. All specimens were 51 mm (2.0 in.) in diameter and 51-mm(2.0-in.) long. Specimens were cored from a solid blockusing a diamond core bit and cut to length by sawing, andends were ground flat. No capp ing material was used in thecompression tests. A steel metal bar, approximately 6.4-mmsquare and 51-mm long ( 0.25 x 0.25 x 2.0-in.), wasmachined to fit on each end of the splitting tensile specimen,as shown in Fig. 4(c).Standard compression tests were performed by applyingthe load along the longitudinal axis of the member [Fig. 4(b)],and splitting tensile tests were performed by rotating thespecimens 90 deg about a transverse specimen axis andapplying a load along a diametrical plane of the specimen, asshown in Fig. 4(c). The dynamic splitting tensile tests in theSHPB were first performed by Ross12 and numericallyanalyzed by Tedesco et al.26Results of wet/dry tests are presented in Fig. 5 and 6 fortension and compression, respectively. In both tension andcompression, the wet and one-half-dry data exhibit greaterstrength increases at higher strain rates than at lower strainrates. In both of these figures, the data is normalized withrespect to the static strength for each type of specimen. In allthe high strain rate work by the authors,8,12,13,15,26 all data isusually normalized with respect to data obtained at lowstrain rate (static test) using the same kind of specimens. Thisresults in the use of the dyna mic increase factor, defined asthe ratio of dynamic strength to static strength. Hopefully,byusing this one may eliminate problems such as differentmaturities relative to cure time and different mix strengths.Also, it is believed that effects of scale due to both specimensize and aggregate may also be minimized by the use of thedynamic increase factor.In all cases the static strengt h of the wet and one-half-dryspecimens was less than that of the static dry tests. However,in the dynamic test the strength of the wet and one-half-dryspecimens was generally greater than the strength of the dryspecimens. The strengths of all the tensile specimens testedare shown as a function of strain rate in Fig. 7.In observing the failure mechanisms of the various tests,one generally finds that in all tensile tests the appearance ofthe fractured surface is the same. The majority of the aggregatesare not fractured but pulled free from the paste. This holdstrue even when comparing dry, one-half-dry, and wet speci-mens to each other as well as for low and high strain ratetests. For compression tests, as the strain rate increases, thenumber of fractured pieces increases and the size of the frac-tured pieces decreases. Fractured surfaces of low strain rate compression specimens are similar to those of the tensilespecimens, but at the higher ra tes, the aggregates begin toshow multiple fractures.Static strength decreases in wet concrete as compared todry concrete have been attributed to the presence of moistureforcing the gel particles apart and reducing the van der Waalsforces.27 These reduced forces are proportional to thespecific surface energy; the criti cal stress for crack formationis then reduced and the overa ll specimen strength is reduced.This explanation appears plausible at low strain rates wherethe crack has time to seek the pa th of least resistance throughthe gel and gel matrix-aggregate interfaces. However, at highstrain rates the presence of the water appears to enhanceconcrete strength. Weerheijm25 attributes the increase of dryconcrete tensile strength at hi gh strain rates to effects ofinertia. The maximum rate at which concrete can fracture islimited, and any effort to force the crack to form at a higherrate simply results in increa sed deformation and strength.The presence of moisture in concrete has a tendency toamplify the inertia effects of the wet specimens, resulting ina strength greater than the dry strength.In addition, the presence of the moisture may also increasematerial stiffness and thereby afford better coupling forwave propagation. The increased stiffness with moisture present is evidenced by the hi gher modulus obtained for wetspecimens, as reported in Re ference 17. Also, the increasedstiffness of wet concrete at hi gh strain rates should result ina higher fracture toughness. A higher fracture toughnessused in Eq. (13) would predict a higher strength for thwet concrete.CONCLUSIONSDry concrete strength is relatively strain rate insensitivebelow the transition strain rates. These transition strain ratesare between 1 and 10/sec for tension, and between 60 and80/sec for compression. Wet co ncrete, however, exhibitssignificant increases in strength at strain rates below as wellas above the transition strain rates for dry concrete. Ananalytical expression that incorporates both crack velocityand fracture toughness as functions of strain rate has beenderived that accurately predicts the tensile fracture strengthof dry concrete. ACKNOWLEDGMENTSThe experimental work for this study was conducted at Wright Laborato-ry Air Base Survivability Section, Tyndall AFB, Florida.REFERENCES1. Watstein, D., “Effect of Straining Rate on the Compressive Strengthand Elastic Properties of Concrete,” ACI J OURNAL, Proceedings V. 24,No. 4, Apr. 1953, pp. 729-744.2. Goldsmith, W.; Kenner, V. H.; and Ricketts, T. E., “DynamicBehavior of Concrete,” Experimental Mechanics, V. 66, 1966, pp. 65-69.3. Melling er, F. M., and Birkimer, D. L., “Measurements of Stress andStrain on Cylindrical Test Specimens of Rock and Concrete under Impact Loading,” Te ch n i c a l R e p o r t 4-46, Department of Army Ohio River Divi- sion Laboratory, Apr. 1966.4. Green, H., “Impact Strength of Concrete,” Institution of Civil Engi-neers, Proceedings, V. 28, 1968, pp. 731-40.5. Hughes, B. P., and Gregory, R., “Impact Strength of Concrete UsingGreen's Ballistic Pendulum,” Institution of Civil Engineers, Proceedings,V. 81, 1968, pp. 731-40.6. Hughes, B. P., and Gregory, R., “Concrete Subjected to High Rates ofLoading in Compression,” Magazine of Concrete Research, V. 24, 1972,pp. 25-36.7. Hughes, B. P., and Watson, A. J., “Compressive Strength and Ulti-mate Strain of C oncrete under Impact Loading,” Magazine of ConcreteResearch , V. 30, 1978, pp. 189-99.8. Malvern, L. E., and Ross, C. A., “Dynamic Response of Concrete andConcrete Structures,” two annual re ports and a final report for AFOSRContract F49620-83-K007, University of Florida, Gainesville, Feb. 1984,Feb. 1985, and May 1986.9. Kormeling, H. A.; Zielinski, A. J.; and Reinhardt, H. W., “Experi-ments on Concrete under Single and Repeated Uniaxial Impact Tensile Loading,” Stevin Laboratory Report 5-80-3, Delft University of Tech-nology, 2nd Printing, 1981.10. Gran, J. K.; Florence, A. L.; and Colton, J. D., “Dynamic TriaxialTests of High-Strength Concrete,” Journal of Engineering Mechanics ,V. 115, 1989, pp. 891-904.11. Suaris, W., and Shah, S. P., “P roperties of Concrete Subjected toImpact Loading,” Journal of Structural Engineering , American Society ofCivil Engineers, V. 109, No. 7, 1983, pp. 1727-1741.12. Ross, C. Engineering and Services Laboratory, Air Force Engineering and Services Center, Tyndall AFB, Florida, 1989.13. Ross, C. A.; Kuennen, S. T.; and Tedesco, J. W., “Experimental andNumerical Analysis of High Strain Rate Concrete Tensile Tests,” Micro- mechanics of Failure of Quasibrittle Materials , S. P. Shah et al., eds.,Elsevier Applied Science, London, 1990, pp. 353-364.14. Sierakowski, R. L., “Dynamic Effects in Concrete Materials,”Application of Fracture Mechan ics to Cementitious Composites , S. P.Shah, ed., Martinus Nijhoff Publishi ng, Dordrecht, Netherlands, 1985.15. Ross, C. A.; Tedesco, J. W.; and Kuennen, S. T., “Effects of StrainRate on Concrete Strength,” ACI Materials Journal, V. 92, No. 1, Jan.-Feb. 1995, pp. 37-47.16. Rossi, P., and Boulay, C., “Etude experimentale du couplagefissuration-viscoité au se in du béton,” Project No. 9, Rapport GRECO, Rheologie des Géomatériaux,1989.17. Reinhardt, H. W.; Rossi, P.; a nd van Mier, J. G. M., “Joint Inves- tigation of Concrete at High Rates of Loading,” Materials and Structures, 23, 1990, pp. 213-216.18. Grad y, E. E., and Lipkin, J., “Criteria for Impulsive Rock Fracture,”Geophysical Research Letters, V. 7, No. 4, Apr. 1980, pp. 255-258.19. Grady, D. E., “Mechanics of Fracture under High Rate Stress Loading,” Sandia National Laboratories Report SAND 82-1148C, 1983. 20. Grady, D. E., and Lipkin, J ., “Mechanisms of Dynamic Fragmen- tation: Factors Governing Fragment Size,” Sandia National Laboratories Report SAND 84-2304C, 1984.21. Broek, D., Elementary Engineering Fracture Mechanics, M. Nijhoff Publishing, Boston, 1982.22. Anderson, T. L., Fracture Mechanics: Fundamentals and Applica-tions , CRC Press, Boca Raton, Florida, 1991.23. John, R., and Shah, S. P., “Effect of High Strain Rate of Loading on Fracture Parameters of Concrete,” Proceedings of Fra cture of Concreteand Rock , Society for Experimental M echanics-Réunion Internationaledes Laboratoires d'Essais et de Recherches sur les Matérieux et les Constructions International Conference, S. P. Shah and S. E. Swartz, eds., Society of Experimental Mechanics, Bethel, Connecticut, 1987, pp. 35-52. 24. Neville, A. M., Properties of Concrete, J. Wiley & Sons, New York, 1973, pp. 313-320.25. Weerheijm, J., “Concrete under Im pact Tensile Loading and Lateral Compression,” PhD dissertation, Delf t Univers ity of Technology, Delft, Netherlands, 1992.26. Tedesco, J. W.; Ross, C. A.; and Brunair, R. M., “Numerical Anal-ysis of Dynamic Split Cylinder Tests,” Computers and Structures, V.32, No. 3-4, 1989, pp. 609-624.27. Wittmann, F. H., “Interaction of Har dened Cement Paste and Water,” Journal of the American Ceramic Society , V. 56, No. 8, Aug. 1972, pp. 409-415.28. John, R., and Shah, S. P., “Fracture of Concrete Subjected to Impact Loading,” Cement, Concrete and Aggregates , CCAGDP V. 8, No. 1, 1986, pp. 24-32. A., “Split-Hopkinson Pressure Bar Tests,” ESL-TR-88-82,。

基于格子链的缩聚反应的动态MonteCarlo模拟

基于格子链的缩聚反应的动态MonteCarlo模拟

基于格子链的缩聚反应的动态Monte Carlo模拟2005年第 63卷第 l3期,1231~1235化学学报ACTA CHIMICA SINICAVO1.63.2005No.13.1231~ 1235研究论文基于格子链的缩聚反应的动态 Monte Carlo模拟吕文琦聚合物分子工程教育部重点实验室丁建东术复旦大学高分子科学系上海 200433摘要采用描述自回避格子链的键长涨落模型,以动态Monte Carlo方法对AB型单体的线型缩聚反应动力学过程进行了模拟.通过该方法可以得到反应过程中链的瞬时构象,还可以得到反应程度、聚合度、分子量分布及其随时间的演化.模拟得到了合理的结果,同时验证了无规线团尺寸与平均链长的标度关系,表明该方法用于研究逐步聚合反应过程是可行的,并且与一般的研究聚合反应的Monte Carlo方法相比,还能够同时得到构象等空间信息.还比较了不同大小的模拟体系所得到的分子量和多分散系数的异同,讨论了有限元胞效应.关键词缩聚反应:动态Monte Carlo方法;格子链;大分子链构象:有限元胞效应Latice Chain Dynamic Monte Carlo Simulation ofPOIyCOndenSatiOn KineticsLO,Wen?Qi DING,Jian?DongKey Laboratory ofMolecular Engineering ofPolymers,Department ofMacromolecular ScienceFudan University,Shanghai 200433Abstract The pO1ycOndensatiOn of linear polymers,based on type AB monomers,was simulated using thedynamic Monte Carlo method with the bond?fluctuation lattice mode1.Besides chain configurations. theconventional parameters such as the extent of reaction,degree of polymerization,molecular weight distribu?tion and their evolution with reaction time can be obtained via statistics.The scaling law of the coil sizeversus chain length during polymerization has also been confirmed.This method is proved to be practical toperform simulation of polymerization kinetics with the advantagethat the spatial inform ation such as con?figuration of self-avoiding chains during polymerization can be obtained. Finite size effect has been furtherexami ned and discussed.Keywords pOlycOndensatiOn;dynami c M onte Carlo simulation;lattice chain;chain configuration;finiteSjze efFect计算机实验目前已经成为化学化工领域的一种重要方法¨J.缩聚反应是重要的高分子合成反应之一,缩聚反应的计算机模拟一般集中在反应动力学及产物的分子量分布问题上.4】,但是,传统的研究聚合反应动力学的 Monte Carlo模拟方法忽略了与空间相关的信息『5’。

Singularity of the density of states in the two-dimensional Hubbard model from finite size

Singularity of the density of states in the two-dimensional Hubbard model from finite size

a r X i v :c o n d -m a t /9503139v 1 27 M a r 1995Singularity of the density of states in the two-dimensional Hubbard model from finitesize scaling of Yang-Lee zerosE.Abraham 1,I.M.Barbour 2,P.H.Cullen 1,E.G.Klepfish 3,E.R.Pike 3and Sarben Sarkar 31Department of Physics,Heriot-Watt University,Edinburgh EH144AS,UK 2Department of Physics,University of Glasgow,Glasgow G128QQ,UK 3Department of Physics,King’s College London,London WC2R 2LS,UK(February 6,2008)A finite size scaling is applied to the Yang-Lee zeros of the grand canonical partition function for the 2-D Hubbard model in the complex chemical potential plane.The logarithmic scaling of the imaginary part of the zeros with the system size indicates a singular dependence of the carrier density on the chemical potential.Our analysis points to a second-order phase transition with critical exponent 12±1transition controlled by the chemical potential.As in order-disorder transitions,one would expect a symmetry breaking signalled by an order parameter.In this model,the particle-hole symmetry is broken by introducing an “external field”which causes the particle density to be-come non-zero.Furthermore,the possibility of the free energy having a singularity at some finite value of the chemical potential is not excluded:in fact it can be a transition indicated by a divergence of the correlation length.A singularity of the free energy at finite “exter-nal field”was found in finite-temperature lattice QCD by using theYang-Leeanalysisforthechiral phase tran-sition [14].A possible scenario for such a transition at finite chemical potential,is one in which the particle den-sity consists of two components derived from the regular and singular parts of the free energy.Since we are dealing with a grand canonical ensemble,the particle number can be calculated for a given chem-ical potential as opposed to constraining the chemical potential by a fixed particle number.Hence the chem-ical potential can be thought of as an external field for exploring the behaviour of the free energy.From the mi-croscopic point of view,the critical values of the chemical potential are associated with singularities of the density of states.Transitions related to the singularity of the density of states are known as Lifshitz transitions [15].In metals these transitions only take place at zero tem-perature,while at finite temperatures the singularities are rounded.However,for a small ratio of temperature to the deviation from the critical values of the chemical potential,the singularity can be traced even at finite tem-perature.Lifshitz transitions may result from topological changes of the Fermi surface,and may occur inside the Brillouin zone as well as on its boundaries [16].In the case of strongly correlated electron systems the shape of the Fermi surface is indeed affected,which in turn may lead to an extension of the Lifshitz-type singularities into the finite-temperature regime.In relating the macroscopic quantity of the carrier den-sity to the density of quasiparticle states,we assumed the validity of a single particle excitation picture.Whether strong correlations completely distort this description is beyond the scope of the current study.However,the iden-tification of the criticality using the Yang-Lee analysis,remains valid even if collective excitations prevail.The paper is organised as follows.In Section 2we out-line the essentials of the computational technique used to simulate the grand canonical partition function and present its expansion as a polynomial in the fugacity vari-able.In Section 3we present the Yang-Lee zeros of the partition function calculated on 62–102lattices and high-light their qualitative differences from the 42lattice.In Section 4we analyse the finite size scaling of the Yang-Lee zeros and compare it to the real-space renormaliza-tion group prediction for a second-order phase transition.Finally,in Section 5we present a summary of our resultsand an outlook for future work.II.SIMULATION ALGORITHM AND FUGACITY EXPANSION OF THE GRAND CANONICALPARTITION FUNCTIONThe model we are studying in this work is a two-dimensional single-band Hubbard HamiltonianˆH=−t <i,j>,σc †i,σc j,σ+U i n i +−12 −µi(n i ++n i −)(1)where the i,j denote the nearest neighbour spatial lat-tice sites,σis the spin degree of freedom and n iσis theelectron number operator c †iσc iσ.The constants t and U correspond to the hopping parameter and the on-site Coulomb repulsion respectively.The chemical potential µis introduced such that µ=0corresponds to half-filling,i.e.the actual chemical potential is shifted from µto µ−U412.(5)This transformation enables one to integrate out the fermionic degrees of freedom and the resulting partition function is written as an ensemble average of a product of two determinantsZ ={s i,l =±1}˜z = {s i,l =±1}det(M +)det(M −)(6)such thatM ±=I +P ± =I +n τ l =1B ±l(7)where the matrices B ±l are defined asB ±l =e −(±dtV )e −dtK e dtµ(8)with V ij =δij s i,l and K ij =1if i,j are nearestneigh-boursand Kij=0otherwise.The matrices in (7)and (8)are of size (n x n y )×(n x n y ),corresponding to the spatial size of the lattice.The expectation value of a physical observable at chemical potential µ,<O >µ,is given by<O >µ=O ˜z (µ){s i,l =±1}˜z (µ,{s i,l })(9)where the sum over the configurations of Ising fields isdenoted by an integral.Since ˜z (µ)is not positive definite for Re(µ)=0we weight the ensemble of configurations by the absolute value of ˜z (µ)at some µ=µ0.Thus<O >µ= O ˜z (µ)˜z (µ)|˜z (µ0)|µ0|˜z (µ0)|µ0(10)The partition function Z (µ)is given byZ (µ)∝˜z (µ)N c˜z (µ0)|˜z (µ0)|×e µβ+e −µβ−e µ0β−e −µ0βn (16)When the average sign is near unity,it is safe to as-sume that the lattice configurations reflect accurately thequantum degrees of freedom.Following Blankenbecler et al.[1]the diagonal matrix elements of the equal-time Green’s operator G ±=(I +P ±)−1accurately describe the fermion density on a given configuration.In this regime the adiabatic approximation,which is the basis of the finite-temperature algorithm,is valid.The situa-tion differs strongly when the average sign becomes small.We are in this case sampling positive and negative ˜z (µ0)configurations with almost equal probability since the ac-ceptance criterion depends only on the absolute value of ˜z (µ0).In the simulations of the HSfields the situation is dif-ferent from the case of fermions interacting with dynam-ical bosonfields presented in Ref.[1].The auxilary HS fields do not have a kinetic energy term in the bosonic action which would suppress their rapidfluctuations and hence recover the adiabaticity.From the previous sim-ulations on a42lattice[3]we know that avoiding the sign problem,by updating at half-filling,results in high uncontrolledfluctuations of the expansion coefficients for the statistical weight,thus severely limiting the range of validity of the expansion.It is therefore important to obtain the partition function for the widest range ofµ0 and observe the persistence of the hierarchy of the ex-pansion coefficients of Z.An error analysis is required to establish the Gaussian distribution of the simulated observables.We present in the following section results of the bootstrap analysis[17]performed on our data for several values ofµ0.III.TEMPERATURE AND LATTICE-SIZEDEPENDENCE OF THE YANG-LEE ZEROS The simulations were performed in the intermediate on-site repulsion regime U=4t forβ=5,6,7.5on lat-tices42,62,82and forβ=5,6on a102lattice.The ex-pansion coefficients given by eqn.(14)are obtained with relatively small errors and exhibit clear Gaussian distri-bution over the ensemble.This behaviour was recorded for a wide range ofµ0which makes our simulations reli-able in spite of the sign problem.In Fig.1(a-c)we present typical distributions of thefirst coefficients correspond-ing to n=1−7in eqn.(14)(normalized with respect to the zeroth power coefficient)forβ=5−7.5for differ-entµ0.The coefficients are obtained using the bootstrap method on over10000configurations forβ=5increasing to over30000forβ=7.5.In spite of different values of the average sign in these simulations,the coefficients of the expansion(16)indicate good correspondence between coefficients obtained with different values of the update chemical potentialµ0:the normalized coefficients taken from differentµ0values and equal power of the expansion variable correspond within the statistical error estimated using the bootstrap analysis.(To compare these coeffi-cients we had to shift the expansion by2coshµ0β.)We also performed a bootstrap analysis of the zeros in theµplane which shows clear Gaussian distribution of their real and imaginary parts(see Fig.2).In addition, we observe overlapping results(i.e.same zeros)obtained with different values ofµ0.The distribution of Yang-Lee zeros in the complexµ-plane is presented in Fig.3(a-c)for the zeros nearest to the real axis.We observe a gradual decrease of the imaginary part as the lattice size increases.The quantitative analysis of this behaviour is discussed in the next section.The critical domain can be identified by the behaviour of the density of Yang-Lee zeros’in the positive half-plane of the fugacity.We expect tofind that this density is tem-perature and volume dependent as the system approaches the phase transition.If the temperature is much higher than the critical temperature,the zeros stay far from the positive real axis as it happens in the high-temperature limit of the one-dimensional Ising model(T c=0)in which,forβ=0,the points of singularity of the free energy lie at fugacity value−1.As the temperature de-creases we expect the zeros to migrate to the positive half-plane with their density,in this region,increasing with the system’s volume.Figures4(a-c)show the number N(θ)of zeros in the sector(0,θ)as a function of the angleθ.The zeros shown in thesefigures are those presented in Fig.3(a-c)in the chemical potential plane with other zeros lying further from the positive real half-axis added in.We included only the zeros having absolute value less than one which we are able to do because if y i is a zero in the fugacity plane,so is1/y i.The errors are shown where they were estimated using the bootstrap analysis(see Fig.2).Forβ=5,even for the largest simulated lattice102, all the zeros are in the negative half-plane.We notice a gradual movement of the pattern of the zeros towards the smallerθvalues with an increasing density of the zeros nearθ=πIV.FINITE SIZE SCALING AND THESINGULARITY OF THE DENSITY OF STATESAs a starting point for thefinite size analysis of theYang-Lee singularities we recall the scaling hypothesis forthe partition function singularities in the critical domain[11].Following this hypothesis,for a change of scale ofthe linear dimension LLL→−1),˜µ=(1−µT cδ(23)Following the real-space renormalization group treatmentof Ref.[11]and assuming that the change of scaleλisa continuous parameter,the exponentαθis related tothe critical exponentνof the correlation length asαθ=1ξ(θλ)=ξ(θ)αθwe obtain ξ∼|θ|−1|θ|ναµ)(26)where θλhas been scaled to ±1and ˜µλexpressed in terms of ˜µand θ.Differentiating this equation with respect to ˜µyields:<n >sing =(−θ)ν(d −αµ)∂F sing (X,Y )ν(d −αµ)singinto the ar-gument Y =˜µαµ(28)which defines the critical exponent 1αµin terms of the scaling exponent αµof the Yang-Lee zeros.Fig.5presents the scaling of the imaginary part of the µzeros for different values of the temperature.The linear regression slope of the logarithm of the imaginary part of the zeros plotted against the logarithm of the inverse lin-ear dimension of the simulation volume,increases when the temperature decreases from β=5to β=6.The re-sults of β=7.5correspond to αµ=1.3within the errors of the zeros as the simulation volume increases from 62to 82.As it is seen from Fig.3,we can trace zeros with similar real part (Re (µ1)≈0.7which is also consistentwith the critical value of the chemical potential given in Ref.[22])as the lattice size increases,which allows us to examine only the scaling of the imaginary part.Table 1presents the values of αµand 1αµδ0.5±0.0560.5±0.21.3±0.3∂µ,as a function ofthe chemical potential on an 82lattice.The location of the peaks of the susceptibility,rounded by the finite size effects,is in good agreement with the distribution of the real part of the Yang-Lee zeros in the complex µ-plane (see Fig.3)which is particularly evident in the β=7.5simulations (Fig.4(c)).The contribution of each zero to the susceptibility can be singled out by expressing the free energy as:F =2n x n yi =1(y −y i )(29)where y is the fugacity variable and y i is the correspond-ing zero of the partition function.The dotted lines on these plots correspond to the contribution of the nearby zeros while the full polynomial contribution is given by the solid lines.We see that the developing singularities are indeed governed by the zeros closest to the real axis.The sharpening of the singularity as the temperature de-creases is also in accordance with the dependence of the distribution of the zeros on the temperature.The singularities of the free energy and its derivative with respect to the chemical potential,can be related to the quasiparticle density of states.To do this we assume that single particle excitations accurately represent the spectrum of the system.The relationship between the average particle density and the density of states ρ(ω)is given by<n >=∞dω1dµ=ρsing (µ)∝1δ−1(32)and hence the rate of divergence of the density of states.As in the case of Lifshitz transitions the singularity of the particle number is rounded at finite temperature.However,for sufficiently low temperatures,the singular-ity of the density of states remains manifest in the free energy,the average particle density,and particle suscep-tibility [15].The regular part of the density of states does not contribute to the criticality,so we can concentrate on the singular part only.Consider a behaviour of the typedensity of states diverging as the−1ρsing(ω)∝(ω−µc)1δ.(33)with the valueδfor the particle number governed by thedivergence of the density of states(at low temperatures)in spite of thefinite-temperature rounding of the singu-larity itself.This rounding of the singularity is indeedreflected in the difference between the values ofαµatβ=5andβ=6.V.DISCUSSION AND OUTLOOKWe note that in ourfinite size scaling analysis we donot include logarithmic corrections.In particular,thesecorrections may prove significant when taking into ac-count the fact that we are dealing with a two-dimensionalsystem in which the pattern of the phase transition islikely to be of Kosterlitz-Thouless type[23].The loga-rithmic corrections to the scaling laws have been provenessential in a recent work of Kenna and Irving[24].In-clusion of these corrections would allow us to obtain thecritical exponents with higher accuracy.However,suchanalysis would require simulations on even larger lattices.The linearfits for the logarithmic scaling and the criti-cal exponents obtained,are to be viewed as approximatevalues reflecting the general behaviour of the Yang-Leezeros as the temperature and lattice size are varied.Al-though the bootstrap analysis provided us with accurateestimates of the statistical error on the values of the ex-pansion coefficients and the Yang-Lee zeros,the smallnumber of zeros obtained with sufficient accuracy doesnot allow us to claim higher precision for the critical ex-ponents on the basis of more elaboratefittings of the scal-ing behaviour.Thefinite-size effects may still be signifi-cant,especially as the simulation temperature decreases,thus affecting the scaling of the Yang-Lee zeros with thesystem rger lattice simulations will therefore berequired for an accurate evaluation of the critical expo-nent for the particle density and the density of states.Nevertheless,the onset of a singularity atfinite temper-ature,and its persistence as the lattice size increases,areevident.The estimate of the critical exponent for the diver-gence rate of the density of states of the quasiparticleexcitation spectrum is particularly relevant to the highT c superconductivity scenario based on the van Hove sin-gularities[25],[26],[27].It is emphasized in Ref.[25]thatthe logarithmic singularity of a two-dimensional electrongas can,due to electronic correlations,turn into a power-law divergence resulting in an extended saddle point atthe lattice momenta(π,0)and(0,π).In the case of the14.I.M.Barbour,A.J.Bell and E.G.Klepfish,Nucl.Phys.B389,285(1993).15.I.M.Lifshitz,JETP38,1569(1960).16.A.A.Abrikosov,Fundamentals of the Theory ofMetals North-Holland(1988).17.P.Hall,The Bootstrap and Edgeworth expansion,Springer(1992).18.S.R.White et al.,Phys.Rev.B40,506(1989).19.J.E.Hirsch,Phys.Rev.B28,4059(1983).20.M.Suzuki,Prog.Theor.Phys.56,1454(1976).21.A.Moreo, D.Scalapino and E.Dagotto,Phys.Rev.B43,11442(1991).22.N.Furukawa and M.Imada,J.Phys.Soc.Japan61,3331(1992).23.J.Kosterlitz and D.Thouless,J.Phys.C6,1181(1973);J.Kosterlitz,J.Phys.C7,1046(1974).24.R.Kenna and A.C.Irving,unpublished.25.K.Gofron et al.,Phys.Rev.Lett.73,3302(1994).26.D.M.Newns,P.C.Pattnaik and C.C.Tsuei,Phys.Rev.B43,3075(1991);D.M.Newns et al.,Phys.Rev.Lett.24,1264(1992);D.M.Newns et al.,Phys.Rev.Lett.73,1264(1994).27.E.Dagotto,A.Nazarenko and A.Moreo,Phys.Rev.Lett.74,310(1995).28.A.A.Abrikosov,J.C.Campuzano and K.Gofron,Physica(Amsterdam)214C,73(1993).29.D.S.Dessau et al.,Phys.Rev.Lett.71,2781(1993);D.M.King et al.,Phys.Rev.Lett.73,3298(1994);P.Aebi et al.,Phys.Rev.Lett.72,2757(1994).30.E.Dagotto, A.Nazarenko and M.Boninsegni,Phys.Rev.Lett.73,728(1994).31.N.Bulut,D.J.Scalapino and S.R.White,Phys.Rev.Lett.73,748(1994).32.S.R.White,Phys.Rev.B44,4670(1991);M.Veki´c and S.R.White,Phys.Rev.B47,1160 (1993).33.C.E.Creffield,E.G.Klepfish,E.R.Pike and SarbenSarkar,unpublished.Figure CaptionsFigure1Bootstrap distribution of normalized coefficients for ex-pansion(14)at different update chemical potentialµ0for an82lattice.The corresponding power of expansion is indicated in the topfigure.(a)β=5,(b)β=6,(c)β=7.5.Figure2Bootstrap distributions for the Yang-Lee zeros in the complexµplane closest to the real axis.(a)102lat-tice atβ=5,(b)102lattice atβ=6,(c)82lattice at β=7.5.Figure3Yang-Lee zeros in the complexµplane closest to the real axis.(a)β=5,(b)β=6,(c)β=7.5.The correspond-ing lattice size is shown in the top right-hand corner. Figure4Angular distribution of the Yang-Lee zeros in the com-plex fugacity plane Error bars are drawn where esti-mated.(a)β=5,(b)β=6,(c)β=7.5.Figure5Scaling of the imaginary part ofµ1(Re(µ1)≈=0.7)as a function of lattice size.αm u indicates the thefit of the logarithmic scaling.Figure6Electronic susceptibility as a function of chemical poten-tial for an82lattice.The solid line represents the con-tribution of all the2n x n y zeros and the dotted line the contribution of the six zeros nearest to the real-µaxis.(a)β=5,(b)β=6,(c)β=7.5.。

多孔介质对流、扩散与导电性

多孔介质对流、扩散与导电性


1
600 500
0.020
Data from Binley et al. (2001) Fit to Eq. [20] Fit by Cassiani et al. (2004)
y = 1.0214x - 0.0007 2 R = 0.9766 0.015
1 3 D p
Data from Katz and Thompson, Advances in Physics, 1987
0 0.4 0.3 0.2 0.1
0.2
0.4
0.6
0.8
1
1.2
1.4
m -1.88
0 -0.1 -0.2 -0.3 -0.4 y = 0.145x - 0.1246 R2 = 0.142
Dp
Effects of Tortuosity, Connectivity
In vicinity of t conductivity must scale as Require both K and dK/d to be continuous at unknown x; this generates prefactor and x.
0.0 0.1 0.2
10-1
10-2
10-3
10-3
b = 0.0 b = 3.0 b = 6.9 b = 10.2 b = 14.3 b = 20.1
0.0 0.1 0.2

0.3
0.4
0.5

0.3
0.4
0.5
Data from Tusheng Ren for silt loams (two adjustable parameters for entire family of curves).

The square-lattice F model revisited a loop-cluster update scaling study

The square-lattice F model revisited a loop-cluster update scaling study

a r X i v :c o n d -m a t /0501222v 2 [c o n d -m a t .s t a t -m e c h ] 20 J u l 2005The square-lattice F model revisited:a loop-cluster update scaling study M Weigel †and W Janke ‡†Department of Physics,University of Waterloo,200University Av W,Waterloo,Ontario,N2L 3G1,Canada ‡Institut f¨u r Theoretische Physik,Universit¨a t Leipzig,Augustusplatz 10/11,04109Leipzig,Germany E-mail:weigel@boromir.uwaterloo.ca ,janke@itp.uni-leipzig.de Abstract.The six-vertex F model on the square lattice constitutes the unique example of an exactly solved model exhibiting an infinite-order phase transition of the Kosterlitz-Thouless type.As one of the few non-trivial exactly solved models,it provides a welcome gauge for new numerical simulation methods and scaling techniques.In view of the notorious problems of clearly resolving the Kosterlitz-Thouless scenario in the two-dimensional XY model numerically,the F model in particular constitutes an instructive reference case for the simulational description of this type of phase transition.We present a loop-cluster update Monte Carlo study of the square-lattice F model,with a focus on the properties not exactly known such as the polarizability or the scaling dimension in the critical phase.For the analysis of the simulation data,finite-size scaling is explicitly derived from the exact solution and plausible assumptions.Guided by the available exact results,the careful inclusion of correction terms in the scaling formulae allows for a reliable determination of the asymptotic behaviour.PACS numbers:75.10.Hk,05.10.Ln,68.35.Rh Submitted to:J.Phys.A:Math.Gen.1.Introduction An ice-type or vertex model was first proposed by Pauling [1]as a model for (type I)water ice.It was known that ice forms a hydrogen-bonded crystal,i.e.,the oxygen atoms are located on a four-valent lattice and the bonding is mediated by one hydrogen atom per bond.Pauling proposed that there be some non-periodicity in the arrangement of the hydrogen bonds in that the hydrogen atoms could be located nearer to one or the other end of the bond.This positioning should satisfy the ice rule ,stating that always two of the bonds are in the “close”position and two are in the “remote”position with respect to the considered oxygen atom.Denoting theposition of the hydrogen atom by a decoration of the bond with an arrow pointing to the closer oxygen,this leads to the arrow configurations depicted in figure 1when for simplicity placing the oxygens on a square lattice instead of the physically realized diamond lattice.Generalizing the resulting six-vertex model for square ice,one assigns energies ǫi ,i =1,...,6,to the vertex configurations depicted in figure 1,resulting in123456Figure1.Allowed arrow configurations for the six-vertex model on the squarelattice,restricted by the ice rule.Boltzmann factorsωi=exp(−βǫi),whereβ=1/k B T is the inverse temperature or coupling.Assuming an overall arrow reversal symmetry(corresponding to the absence of an external electricfield),one abbreviates a=ω1=ω2,b=ω3=ω4 and c=ω5=ω6.Then,the original ice model corresponds to the choiceǫi=0, i=1,...,6,whereas another especially symmetric version assumesǫa=ǫb=1,ǫc=0resp.a=b=e−β,c=1,(1) which is known as the F model of anti-ferroelectrics[2],since due to the choice of weights the vertex configurations5and6will dominate for low temperatures,resulting in a ground-state of staggered,anti-ferroelectric order as depicted infigure2.The six-vertex model as well as the more general eight-vertex models,obtained by including sink and source vertices with all four arrows pointing in and out,respectively, have been exactly solved in zerofield using transfer matrix techniques,see reference[3]. They exhibit rich phase diagrams featuringfirst-order and continuous phase transitions as well as multi-critical points.In particular,the six-vertex F model undergoes an infinite-order transition of the Berezinskii-Kosterlitz-Thouless(BKT)type to an anti-ferroelectrically ordered phase and the scaling behaviour of the basic thermodynamic quantities can be extracted from the closed-form solution.Since there is no solution of the model in a(staggered)field,however,information about properties related to the polarization is incomplete.The same is true for the correlation function,which can only be evaluated at the so-called free-fermion point of the model[3,4](the correlation length,however,is exactly known for all temperatures,see below).Also, since the solution was obtained in the thermodynamic limit,information aboutfinite-size scaling(FSS)is not exact,but must be deduced from scaling arguments.Apart from its prominent position as a non-trivial solvable model of statistical mechanics, the F model has enjoyed sustained interest due to its equivalence to the BCSOS surface model[5],and hence several dynamical generalizations of the six-vertex model have been considered[6].A six-vertex model with so-called domain-wall boundary conditions has recently attracted considerable interest and found numerous applications in counting problems,the quantum inverse scattering method etc.[7].The Berezinskii-Kosterlitz-Thouless[8]scenario of an infinite-order phase transition induced by the unbinding of vortex pairs in the two-dimensional XY model has been found exceptionally hard to confirm numerically[9,10,11,12].This is partially due to the nature of the infinite-order phase transition itself,which is not easy to distinguish from afinite-order transition numerically,and the presence of a critical phase,which render many of the standard FSS techniques less useful.The main trouble,however,is caused by the presence of logarithmic corrections,expected to be present on general grounds for a theory with central charge c=1[13]and explicitlyFigure 2.Cutout of one of the two anti-ferroelectrically ordered groundstates of the square-lattice F model.The state consists of vertices5and6atequal proportions.The dashed lines indicate one of two tilted sub-lattices offerroelectrical order.found from the BKT theory of the model[14].While a numerical confirmation of the leading scaling behaviour of the BKT transition has been achieved in the past decade or so,the resolution of the logarithmic scaling corrections is still at the forefront of problems amenable to numerical investigation today[15,16,17].From duality arguments and mapping to Coulomb gas systems,the F model is known to be asymptotically equivalent to the two-dimensional XY model at criticality. Thus,apart from being an interesting subject in its own right,a detailed analysis of the thermal and FSS properties of the six-vertex F model in the critical phase and at its BKT point serves as a guideline for simulations of the XY model case.Guidance is been given here through the fact that the exact solution of the F model yields the leading singularities including the correction terms explicitly,and,most notably for numerical purposes,the exact critical coupling of the model.Uncertainties occurring in analyses of the XY model such as systematic errors in the determination of the transition point or the effect of neglected higher-order correction terms can be studied rather explicitly for the F model.Finally,when it is found here that one has to consider large system sizes and proceed carefully when including correction terms into thefits,this situation should also be put into relation with the case of an F model placed on an annealed ensemble of random lattices considered recently[18].Guided by the present investigation,this case has to be analyzed even more carefully due to an additional fractality of the lattices,which reduces the effective linear extent of the amenable lattice sizes,thus increasingfinite-size effects even further.The other paradigm example of an exactly solved non-trivial model of statistical mechanics,the two-dimensional Ising model,has served as a benchmark system and playground for new ideas in the theory of critical phenomena as well as for new algorithms in computer simulations in an overwhelming number of studies,and almost all of its aspects have been investigated(but not necessarily understood).In contrast, for the case of vertex models only rather recently efficient cluster-update Monte Carlo algorithms have been developed[19,20,21],mainly with the mapping of vertex models on quantum chains in mind,and some simulations of special aspects of the six-vertex model,such as dynamical critical exponents of the considered algorithms[20,22], properties of the equivalent surface models[23],matching of renormalization-groupflows with the XY model[24],or the case of domain-wall boundary conditions[21] have been analyzed.A systematic thermal and FSS study of the F model in the critical high-temperature phase,at the critical point and its low-temperature vicinity including the analysis of the logarithmic correction terms,however,is to our best knowledge lacking so far.The rest of the paper is organized as follows.In section2we outline the extent of exact knowledge about the phase diagram and the occurring transitions of the six-vertex model and the F model in particular and give an overview over scaling at a BKT point in general.After a short description of the simulational setup used,section 3contains a report of the analysis of the simulation data,comprising the FSS analyses of the critical-point thermodynamic properties(where the corresponding FSS relations are explicitly derived from the closed-form solution),an investigation of the behaviour in the critical phase as well as a thermal scaling analysis in the low-temperature phase of the model.Finally,section4contains our conclusions.2.Analytical Results2.1.Exact solution and phase diagramThe square-lattice,zero-field six-vertex model has been solved exactly in the thermodynamic limit by means of the Bethe ansatz by Lieb[25]and Sutherland[26]. The analytic structure of the free energy is most conveniently parametrized in terms of the reduced couplinga2+b2−c2∆=in terms of the re-scaled weights a/c and b/c.Phase boundaries are indicated bysolid lines.The dashed line denotes the parameter range of the F model.Thedotted line corresponds to the free-fermion line∆=0.While the ferroelectrically ordered phases I and II exhibit a plain polarization, which can be used as an order parameter for the corresponding transition,the anti-ferroelectric order of phase IV is accompanied by a staggered polarization with respect to a sub-lattice decomposition of the square lattice.This is equivalent to a mutually inverse plain polarization on two tilted,square sub-lattices as indicated infigure2. An order parameter for the corresponding transition can be defined by introducing overlap variablesσi for each vertex i such that[3],σi=v i∗v0i,(3) where v0i denotes the anti-ferroelectric ground-state configuration depicted infigure2 and the product“∗”denotes the overlap given byv∗v′≡4k=1A k(v)A k(v′),(4)where k numbers the four edges around each vertex and A k(v)should be+1or−1depending on whether the corresponding arrow of v points out of the vertex or into it.The thus defined spontaneous staggered polarization P0≡ σi /2= σ /2constitutes an order parameter for the antiferroelectric transition III→IV.As indicated above,the F model orders anti-ferroelectrically at∆=−1, corresponding to a critical couplingβc=ln2.From the exact solution[3,25,26], the model’s asymptotic free energy per site in the low-temperature phaseβ>βc can be written asf low(λ)=β−λm cosh(mλ),(5)whereλ=acosh[14µ ∞0d t cosh t−1,(6)where µ=arccos[11+x 2m −1 2=∞ m =11−y 2m −1λ 1/2∞n =1y (n −1The specific heat,on the other hand,is only very weakly singular,behaving as (omitting a regular background contribution)C v∼ξ−2.(13) Finite-size scaling analyses of the BKT transition are hampered by the occurring essential singularities and the presence of a critical phase.As a consequence of the latter,magnetic observables such as the susceptibility do not exhibit maxima in the vicinity of the critical point,which otherwise could be used for an estimation of the transition temperature fromfinite systems.For the same reason,also the Binder parameter requires a more careful treatment than at a standard second-order phase transition[29,30].Nevertheless,the general arguments forfinite-size shifting and rounding remain valid,such that suitably defined pseudo-critical points T∗(L)for systems with linear extent L scale to the critical point T c as[31][T∗(L)−T c]/T c∼(ln L)−1/ρ,(14) cf.equation(11),since sufficiently close to the critical point the growth of the correlation length becomes limited by the linear extent L of the system. Correspondingly,ξcan be replaced by L(neglecting corrections to scaling for the time being)to yield the FSS lawχ(T c,L)∼Lγ/ν=L2−ηc,(15) which forηc=1/4predicts a rather strong divergence.Onfinite lattices,the specific heat is found to exhibit a smooth peak,which is however considerably shifted away from the critical point into the high-temperature phase and does not scale as the lattice size is increased[31].Thus,with the main strengths of FSS being not exploitable for the BKT phase transition,the focus of numerical analyses of the XY and related models has been on thermal scaling,see,e.g.,references[9,10].In addition,renormalization group analyses predict logarithmic corrections to the leading scaling behaviour[14],as expected for a theory of central charge c=1,which have been found exceptionally hard to reproduce numerically due to the presence of higher order corrections of comparable magnitude(for the accessible lattice sizes)[11,12,17].From the exact solution of the square-lattice F model,equations(5)–(8),one extracts the asymptotic behaviour in the vicinity of the critical pointβc=ln2. Approaching the critical point from the low-temperature side,λ↓0,the singular part of the free energy density(5)and the correlation length(7)behave asf sing(λ)∼4k B T c exp(−π2/λ),(16)ξ−1(λ)∼4exp(−π2/2λ).Sinceλgoes asλ∼(−t)1/2for t↑0,this exactly corresponds to the essential singularity described above for the BKT transition of the two-dimensional XY model withρ=1/2.The specific heat has the weakly singular contribution C v∼ξ−2as expected.Concerning properties related to the order parameter,the situation for the F model is somewhat different from that of the XY model.The order parameter(8)is non-vanishing forfinite temperatures in the ordered phase§.Thus,the corresponding staggered anti-ferroelectric polarizabilityχshows a clear peak in the vicinity of the critical point forfinite lattices.However,in the limit L→∞the polarizability diverges throughout the whole critical high-temperature phase.Note that compared to the XY §Note that the Mermin-Wagner-Hohenberg theorem[28]does not apply to the F model with its discrete symmetry.model the rˆo les of high-and low-temperature phases are exchanged in this respect,asexpected from duality[32].The spontaneous polarization(8)scales asP0(λ)∼λ−1exp(−π2/4λ)∼ξ−1/2lnξ(17) asλ↓0,implyingβ/ν=1/2.Assuming the Widom-Fisher scaling relationα+2β+γ=2to be valid ,from equations(16)and(17)Baxter conjectured the following scaling of the zero-field staggered polarizability[27],χ(λ)∼λ−2exp(π2/2λ)∼ξ(lnξ)2,(18) which impliesγ/ν=2−ηc=1,obviously different from the XY model resultηc=1/4. Since the whole high-temperature phase is critical,scaling of the polarizability isexpected throughout this phase.In fact,for the free-fermion case∆=0orβ=βf≡(ln2)/2,which is exactly solvable in a staggeredfield[4],a logarithmic divergence of the polarizability is found,implying2−ηf=0.More recently,thebehaviour ofηin the critical phase of the F model has been conjectured from scattering methods to follow the form[23,34,35]η(∆)=π/arccos∆.(19) This is in agreement with the exact results for the critical F model at∆=−1and the free-fermion case at∆=0.Additionally,the pure ice model at∆=1/2is known to have a“dipolar”correlation function withη=3as predicted by(19)[34,36]. Note that,since the dual relation to the XY model is only valid at criticality and the XY model magnetization is not equivalent to the polarizability of the F model, this result is not simply related to the exponentηof the XY model in its critical low-temperature phase,which actually decreases as one moves into the critical phase, see,e.g.,references[14,37].The common occurrence of a BKT type phase transition for the XY and F models is no coincidence.In fact,it can be shown that the critical points of both models are asymptotically dual to each other[32].This can be seen by noting that the Villain representation of the XY model[38]is dually equivalent to a model of the solid-on-solid(SOS)type known as the discrete Gaussian model[39],which in turn,as typical for SOS models,can be mapped onto the Coulomb gas[40].The F model,on the other hand,also has a height-model representation known as the BCSOS(body-centered SOS)model[5],which is itself asymptotically equivalent to the Coulomb gas. Alternatively,the stated equivalence can be seen from the loop representation of the O(n)vector model[41],which for the critical O(2)model yields a close-packed loop ensemble equivalent to that of the loop representation of the critical F model[42]. The apparent discrepancy regarding the magnetic exponentsβ/νandγ/νbetween the XY and F models,on the other hand,is not an indicator of different universality classes of the models,but reflects the fact that the F model staggered polarizability is not equivalent to the magnetic susceptibility of the XY model.Although the BKT transition is characterized by essential singularities and thus the conventional critical exponents are meaningless,one can re-define them by considering scaling as a function of the correlation lengthξinstead of the reduced temperature t[33].The exponentsα,βandγused here and in the following should be understood in this sense.The exponentρ,however,has its special meaning defined by(11).3.A Loop-Cluster Update Scaling Study3.1.Simulation SetupFor an analysis of the six-vertex F model via Monte Carlo simulations,a suitable simulation update scheme has to be devised.Since the focus here lies on the investigation of the vicinity of the BKT transition and the critical phase of the model, all local updates will suffer from the severe critical slowing down with dynamical critical exponent z≈2expected at or close to criticality.Fortunately,a fully-fledged framework of cluster algorithms has been constructed for the simulation of the six-and eight-vertex models,mainly motivated by their equivalence with the Trotter-Suzuki decomposition of quantum spin chains.Here,we apply the so-called loop-cluster algorithm[43],which operates on a representation of the vertex model by polygons consisting of the lattice edges and induced by a stochastic breakup of the lattice vertices,for details see reference[43].For the case of the F model at criticality, a reduction of critical slowing down to z=0.71(5)has been reported[44].Simulations were performed for square lattices with periodic boundary conditions, measurements were taken after each multi-cluster loop-update step due to the small autocorrelation times observed.To enable a proper FSS analysis,for the investigation of the BKT point two main series of simulations were performed;one around the peak locations of the staggered anti-ferroelectric polarizability for sizes L=16,24, 32,46,64,92,128,182,and256and another at the asymptotic critical coupling βc=ln2=0.6931...with additional lattice sizes of L=364,512,726and1024.To examine the behaviour in the critical phase,additional series of simulations have been performed atfixed temperaturesβ<βc with lattice sizes identical to those atβc.Per simulation,after equilibration a total of between1×105and2×105measurements were taken.3.2.Results of the Finite-Size Scaling Analysis3.2.1.Non-scaling of the specific heat.The specific heat is defined byC v=β2[ E2 − E 2]/L2,(20) with the internal energy E of a vertex configuration,E= i E(v i),E(v i)∈{ǫ1,...,ǫ6},(21)where v i denotes the configuration of vertex i of the lattice.It exhibits a broad peak shifted away from the critical point into the low-temperature phase[45]¶.The essential singularity predicted by equation(13)cannot in general be resolved,since it is covered by the presence of non-singular background terms.Thus,the non-scaling of a broad specific-heat peak(together with a scaling of the susceptibility or polarizability to be considered below)is commonly taken as afirst good indicator for a phase transition to be of the BKT type[31].Indeed,this is what is found from the simulation data as is shown infigure4.No scaling is visible,apart from very minor deviations for the smallest lattice sizes and close to criticality.All data points collapse onto a single curve,which is identical to the exact asymptotic behaviour of C v extracted from the free energy density of equations(5)and(6)as displayed infigure4for comparison.¶Recall that the specific heat of the2D XY model exhibits a peak in the high-temperature phase [10],as expected from duality.Figure 4.Non-scaling of the specific heat C v of the square-lattice F model.For clarity,simulation results from only three lattice sizes are shown.The solidline denotes the exact asymptotic result found from the free energy density ofequations(5)and(6).The dashed vertical line marks the infinite-volume criticalpointβc=ln2=0.6931....In particular,at the critical pointβc=ln2wefind for the internal energy U= E and the specific heat C v for the10242lattice,U(βc)=0.333335(4),(22)C v(βc)=0.3005(15),in perfect agreement with the exact results U(βc)=1/3and C v(βc)=28(ln2)2/45≈0.2989[45].3.2.2.The critical coupling.For an independent determination of the critical couplingβc from the simulation data,we exploit the fact that the maxima of the staggered polarizability forfinite lattices should be shifted away from the critical point according to the scaling relation(14).Fromfinite-lattice simulations the polarization is determined by breaking the symmetry explicitly,i.e.,if one definesΣ= iσi,the spontaneous polarization is measured as P0= |Σ| and the polarizability is estimated byχ=[ Σ2 − |Σ| 2]/L2.(23) The peak locations ofχ(β)were determined from simulations at nearby couplingsβby means of the reweighting technique[46].The phenomenological theory of FSS[31] implies that the polarizabilityχfor afinite lattice can be expressed asχ(β,L)=Lγ/νX[L/ξ(β,∞)],(24) whereξ(β,∞)=ξ(β,L=∞)denotes the correlation length of the infinite system and X is an analytic scaling function(here,we omit additional irrelevant scalingfields representing corrections to scaling).Now,the maxima ofχ(β,L)correspond to themaximum of X and thus all must occur at the same value of the argument L/ξ(β,∞)(provided X only has one maximum),L2 ln 148ξ−3+O (ξ−5) −1,(26)and β(λ)expands around the critical point λ=0asβ=ln 2+1192λ4+O (λ6).(27)To leading order in both expansions one thus has via equation(25)β∗(L )=βc +A β(ln 4ξ)−2=βc +A β(ln 4κL )−2,(28)where A β=(π2/4√3)2.We tested fits of this expected asymptotic form to thesimulation data for the toy model of an analytically generated series of pseudo-critical points β∗(L )defined by equation (25)and the exact form of the correlation length(7)with κ=1and L =16,24,...,256as in the simulations,while taking βc ,κand the amplitudes A βand C βas fit parameters.Already without the correction,i.e.,enforcing C β=0,the critical coupling is reasonably reproduced as βc =0.692;the presence of neglected corrections shows up,however,in a fit result κ≈1.3.Lifting the constraint on the amplitude C β,one arrives at βc =0.6931and κ=1.05,indicating perfect agreement with the input data on the level of accuracy to be expected from the simulations.Considering real simulation data,one might actually want to replace ξ(β,∞)in(25)by the finite-size expression ξ(β,L )(or,equivalently,allow the constant κto depend on system size),introducing additional corrections not explicitly included in the scaling form (24)+.Exact expressions for the finite-size behaviour of ξare not available ∗,however,in view of the scaling forms (17)and (18)and the experience with various models with logarithmic scaling corrections,it seems reasonable to make the following general ansatz,ξ[β∗(L ),L ]=κL [1+A ξ(ln 4κL )ωξ](30)+Note that for the case of models with multiplicative logarithmic corrections,the replacement ξ(β,∞)→ξ(β,L )on the r.h.s.of (24)has been suggested as the proper way of describing FSS in the first place [47].∗Note,however,that exact expression are available for the finite-size correlation length of the XY model on the strip geometry,cf.reference [48].Table 1.Parameters of least-squares fits of the functional form(31)tothesimulationestimatesforthepeak locations of the staggered polarizability.No correction terms were taken into account,i.e.,B β=C β=0were held fixedthroughout.Q denotes the quality-of-fit parameter.160.73822(48)0.3547(62)0.00240.73270(59)0.4533(87)0.00320.73033(74)0.5018(126)0.00460.72635(110)0.5912(223)0.46640.72409(172)0.6453(385)0.88920.72322(261)0.6668(624)0.781280.72077(463)0.7307(1173)0.79ln L ,ωξ>0.(32)An indirect determination of the finite-size correlation length to be discussed below in section 3.2.3strongly hints at the presence of only additive logarithmic corrections in (30),implying ωξ<0,but in some cases both possibilities will be considered here to illustrate the fact that a numerical discrimination between similar forms of the corrections is not at all easily possible [note that due to the extremely slow variation of the log-log term it might effectively be considered constant for the range of lattice sizes considered,which would render the form (32)equivalent to the ansatz (31)].The determined peak locations of the polarizability together with an example fit of the functional form (31)with omitted corrections,i.e.,for B β=C β=0,to the data in the range L =L min =92up to L =256,are shown in figure 5.The fit parameters of such fits,successively omitting points from the low-L side,are compiled in table 1.The strong deviations of the data from the form with B β=C β=0corresponding to a straight line in the chosen scaling of the axes are apparent from figure pared to the exact transition point βc =ln 2,the estimates of βc from these fits are clearly too large,dropping only very slowly as points from the small-L side of the list are successively omitted,cf.table 1.One might attempt to extrapolateFigure5.Peak positions of the staggered anti-ferroelectric polarizability of theF model from simulations as a function of lattice size.The lines showfits of theform(31)to the data.The dashed line corresponds to an uncorrectedfit withBβ=Cβ=0,starting from L min=92.The solid curve corresponds to afit withcorrections and L min=32.these results towards L min→∞using the scaling form(31);since the individual data are highly correlated,however,this would introduce a strong bias.Instead,we directly use the higher-order logarithmic corrections in thefitting procedure.Note that this effect of strong scaling corrections here occurs for rather large lattices,where for a usual continuous phase transition without logarithmic corrections the presence of scaling corrections usually would not be much of an issue for the determination of the leading scaling behaviour.Relaxing the constraints on Bβonly or on both parameters, Bβand Cβ,we arrive at thefit results compiled in table2.It is apparent that the complexity of the completely unconstrainedfit type is at the verge of exceeding the available statistical accuracy of the data,such that competing local minima of theχ2 distribution exist,which result in a rather discontinuous evolution of the amplitudes Aβ,Bβand Cβas the lower-end cut-offL min is increased.This functional formfits the data very well,however,and the estimates forβc are all in agreement with the asymptotic valueβc=0.6931...in terms of the statistical errors.It seems clear that the remaining vague tendency of thefits to yieldβc slightly above its asymptotic value could in principle be removed by including further correction terms for the case of extremely accurate data.Adding a(ln L)−5term in(31),for instance,yields βc=0.674(45),Q=0.72when including all data points.Using the alternativefit form(32)valid forωξ>0,on the other hand,results in fits quite similar to those obtained from the ansatz(31)with both correction terms present.This relates back to the remark concerning the extremely slow variation of the log-log term,which makes it plausible that the considered correction terms can be effectively interchanged.No specific drift is observed on increasing the cutoffL min. For L min=64we arrive atβc=0.695(40),Q=0.92.Hence,from the scaling of the pseudo-critical(inverse)temperaturesβ∗(L),a conclusion about the sign ofωξcan hardly be drawn.。

Monte Carlo simulations of the critical properties of the restricted primitive model

Monte Carlo simulations of the critical properties of the restricted primitive model

a r X i v :c o n d -m a t /0409356v 2 [c o n d -m a t .s t a t -m e c h ] 24 S e p 2004Monte Carlo simulations of the critical properties of therestricted primitive modelJean-Michel Caillol Laboratoire de Physique Th´e orique UMR 8267,Bˆa t.210Universit´e de Paris-Sud 91405Orsay Cedex,France ∗(Dated:February 2,2008)Abstract Recent Monte Carlo simulations of the critical point of the restricted primitive model for ionic solutions are reported.Only the continuum version of the model is considered.A finite size scaling analysis based in the Bruce-Wilding procedure gives critical exponents in agreement with those of the three-dimensional Ising universality class.An anomaly in the scaling of the specific heat with system size is pointed out.PACS numbers:1I.INTRODUCTIONThe primitive model(PM)for electrolytes,molten salts,collo¨ids,etc is a mixture of M species of charged hard spheres living either on a lattice or within a continuous volume of real space.In this paper we shall focus only on the off-lattice version of the model.The simplest version of the PM consists of a binary mixture(i.e.M=2)of positive and negative charged hard spheres±q all with the same diameterσ.Under this form the model which is thought to be the prototype of many ionicfluids has been christened the restricted primitive model(RPM).A thermodynamic state of the RPM is entirely specified by a reduced density ρ∗=Nσ3/V(N number of ions,V volume)and a reduced temperature T∗=kTσ/q2(k Boltzmann’s constant).The RPM undergoes a liquid-vapor transition which has been studied extensively these last past years by means of Monte Carlo(MC)simulations and various theoretical ap-proaches.In particular the behavior of the system at its critical point(CP)has been the subject of a huge amount of numerical and theoretical studies.The question is obviously of great importance since it is reasonable to assume that real electrolytes-or at last a large class of them-and the RPM belong to the same universality class which dictates a similar critical behavior.It is perhaps the right place to note that the main feature of ionic solutions is that the pair potential between two ions i and j at a distance r ij=|r i−r j|which reads asq i q jv(r ij)=support this claim.Most numerical studies of the CP of the off-lattice version of the RPM were performed by the Orsay group and I would like to review our contributions towards a better understanding of the critical properties of this model in the lines below.II.A BRIEF HISTORICAL SUR VEYQuite generally,a single componentfluid will undergo a liquid vapor transition if the pair potential which is assumed to represent the molecular interactions is(sufficiently)attractive at large distances.From this point of view the situation is not so clear in the case of the RPM(cf eq.(1.1))and the very existence of the transition is not guaranteed.Several studies were necessary to clarify this point and a brief historical survey is worthwhile.Thefirst evidence that the RPM actually undergoes a liquid-vapor transition can be tracked back to two papers of Chasovkikh and Vorontsov-Vel’Yaminov(CVVY)published as soon as in19766,7.These authors performed isobaric MC simulations and found a tran-sition with a CP located at T∗c=0.095,ρ∗c=0.24.Several years after(in1991)Valleau studied three isotherms of the RPM with his method of the density scaling MC and ob-tained a different location for the CP,namely T∗c=0.07,ρ∗c=0.078.Subsequently(in1992) Panagiotopoulos9obtained still different results,i.e.T∗c=0.056,ρ∗c=0.04,by performing MC simulations in the Gibbs ensemble(GE),at the moment a powerful new method of simulation which he had invented a little bit earlier10.Subsequent GE simulations using an improved biased MC sampling11yielded Panagiotopoulos and Orkoulas to the new estimate T∗c=0.053,ρ∗c=0.025.Finally,making use myself of the Gibbs ensemble combined with the use of hyperspherical geometries I obtained rather T∗c=0.057,ρ∗c=0.0412,13.Commenting on this striking dispersion of the MC data Prof M.Fisher talked once of the ”sad street of numerical simulations”.This was in1999,at the SCCS conference,St Malo, France and,at this point of the story,I must agree with him retrospectively.However many advances have been done since.Before giving an account of these new achievements some comments are in order.•(i)All the MC studies confirm the existence of a liquid vapor transition for the RPM.It seems to take place at unusually low densities and temperatures.Caillol and Weis give further support for such a low critical temperature14.Moreover it turns out that the coexistence curve is very dissymmetric9,11,12.•(ii)The MC simulation of ionic systems is a numerical challenge due to the long range of Coulomb potential.In order to deal with this,some caution is needed.Thus,in the case of MC simulations performed in a cubic box with periodic boundary conditions (PBC),one must use Ewald potentials in order to obtain the correct physics15,16,17.The point is that the Ewald potential is the solution of Poisson equation in a cubico-periodical geometry17and many properties of ionicfluids(electro-neutrality,screening, etc)are a consequence of this fact.In their MC simulations CVVY and Valleau considered truncated Coulomb potentials and very small samples of N=32particles which yields quantitatively wrong results.By contrast the data of Panagiotopoulos et al.9,11are more reliable since Ewald sums have been used.The same remark apply to my simulations which were performed on a4D sphere(a hypersphere for short)by considering interactions obtained by solving Poisson equation in this geometry.This alternative method of simulation is therefore also indisputably correct,moreover it is much more efficient.The rough agreement observed between the simulations of refs.9,11and12both involving the same number of ions,i.e.N=512,is therefore not fortuitous.•(iii)None of the above mentioned studies took correctly into accountfinite size ef-fects which are of an overwhelming importance near a CP.These effects affect the behavior offinite systems as soon as the correlation length of the critical densityfluc-tuations is of the same order of magnitude than the size of the simulation box.In the simulations9,11,12some”apparent”critical temperature T∗c has been measured which could be very different from its infinite volume limit T∗c(∞).In order to extract from MC simulations the critical behavior of the RPM in the thermo-dynamic limit(i.e.the critical exponents)and also the infinite volume limit of T∗c andρ∗c it is necessary to perform an analysis of the MC data in the framework of thefinite size scaling (fss)theory which is part of the renormalization group(RG)theory18,19.In this approach one needs to work in the Grand Canonical(GC)ensemble rather than in the Gibbs ensemble which is ill adapted for a fss analysis.Subsequent MC simulations on the RPM were thus all performed in this ensemble.Panagiotopoulos and coworkers turned their attention to the lattice version of the RPM whereas the Orsay group continued to work on its off-lattice version.III.FINITE SIZE SCALING ANALYSIS OF MC DATAA.Scalingfields and operatorsStarting with the seminal work of Bruce and Wilding(BW)20,21,22simulation results for the critical behavior offluids have customarily been analyzed along the lines of the so-called revised scaling theory of Rehr and Mermin23.In this approach onefirst defines scalingfields and operators aimed at restoring the particle-hole symmetry and therefore to map the the fluid onto a magnetic sytem with Ising-like symmetry.The two relevant scalingfields h(the strong orderingfield)andτ(the weak thermal field)are assumed to be linear combinations of deviations from their critical values of the chemical potentialµand the inverse temperatureβ=1/T(reduced values are assumed henceforward).One thus hash=µ−µc+r(β−βc)τ=βc−β+s(µ−µc),(3.1)where r and s are thefield mixing parameters which define the mapping.Of course relations (3.1)are valid only in the vicinity of the CP.The conjugate scaling operators M and E are then defined as<M>=1∂hlnΞ=1V∂1−sr(<u>−r<ρ>),(3.2)whereΞis the GC partition function of the RPM,ρthe total number density,and u the internal energy per unit volume.Brackets<...>denote GC averages.M is the order parameter(magnetization)of the magnetic system associated with thefluid and E its mag-netic energy.E should be invariant under the transformations(M,h)→(−M,−h)for appropriate values of s and r.In this framework the coexistence curve is therefore defined by the eq.h=0.The revised scaling of Rehr and Mermin implies the analyticity of the coexistence chemical potentialµ(T)at T∗c.Although this is the case for some peculiar lattice gas models with ”hidden”symmetries there is no reason that in general,forfluid systemsµ(T)should lack asingularity as recognized already by Rehr and Mermin23and emphasized more recently by Fisher and co-workers40,41,42.B.The scaling hypothesisA central role in the subsequent fss analysis is played by the GC joint distribution P L(M,E)∝P L(ρ,u)for the scaling operators M and E.Following BW20,21,22we will assume that,in the immediate vicinity of the CP,P L(M,E)obeys to the following scaling law:P L(M,E)=a−1M a−1EL d−yτL d−y h P(a−1ML d−y hδM,...a−1EL d−yτδE,a M L y h h,a E L yττ,a i L y i,...),(3.3) where L are the linear dimensions of the system(taken as V1/3,where V is the volume of the simulation box,either a cube or a hypersphere).I have denoted byδM≡M−<M>c andδE≡E−<E>c the deviations of the scaling operators from their value at criticality. The cornerstone of this scaling hypothesis is that the function P which enters eq.(3.3)is universal in the sense that it depends only upon the universality class of the model and of the type of geometry considered.The constants a M,a E,and a i are system dependent constants which are defined in such a way that P has unit variance.Finally,the renormalization exponents y h,yτ,and y i which enter eq.(3.3)are defined asy h=d−β/νyτ=1/νy i=−θ/ν(3.4) in terms of the usual critical exponents:•βexponent of the orderingfield,i.e.<δM>∼|τ|βfor T∗<T∗c at h=0•νexponent of the correlation length,i.e.ξ∼|τ|−ν•θWegner’s correction-to-scaling exponent(first irrelevant exponent).The scaling hypothesis(3.3)was established on a solid RG basis for Ising-like systems24and received substantial supports from MC studies25.We stress once again that the coexistence curve is determined in this approach by the condition h=0and that,at coexistence,the order parameter distribution P L(M)should be an even function of M.In practice this symmetry requirement can be satisfied by tuning the two parameters(µ,s)at a givenβ. We now concentrate our attention on the scaling behavior of the histogram P L(M).C.The matching procedureIntegrating both sides of eq.(3.3)over E onefinds that,along the coexistence line h=0 one hasP L(M)=a−1M L d−y h P(a−1ML d−y hδM,a E L yττ,a i L y i),(3.5)where,in the r.h.s.the dependence of the universal function P upon h has been discardedfor clarity.Let us define now x=a−1ML d−y hδM,then,assumingτ∼0and L∼∞a Taylor expansion of eq.(3.5)yieldsP L(M)=a−1ML d−y h P∗(x)+a E L yττ P∗1(x)+a′2E L2yττ2 P∗2(x)+...+a i L y i P∗3(x)+... ,(3.6) where the various P∗entering the r.h.s.are universal functions.Note that,for L=∞the normalized orderingfield distribution P L(M)collapses onto an universal function P∗(x)at τ=0.For Lfinite but large P L(M)collapses approximately onto P∗(x)at some apparent τL∝L−yτ+y i.Since for h=0one hasτ∝β−βc then the matching of the histogram P L(M) onto the universal function P∗(x)should occur at some apparent temperature T∗c(L)scaling with system size asT∗c(∞)−T∗c(L)∝L−(θ+1)/ν+....,(3.7) where T∗c(∞)denotes the infinite volume limit of the critical temperature.D.Technical detailsTo assess the critical behavior and the critical parameters of the system,we need,in a first step,to locate the coexistence curve h=0.At a given temperatureβclose toβc theordering distribution function P L(M)depends solely on the chemical potentialµand the mixing parameter s.At coexistence,the value of(µ,s)can be obtained unambiguously by requiring that P L(M)is symmetric in M−<M>22.Tuning at willµand s at givenβrequires to know the joint histogram P L(M,E)∝P L(ρ,u)for a continuous set of values of µat a givenβ.Moreover,since this analysis must be performed at differentβone needs in fact to know P L(M,E)for a continuous set of values of(β,µ)in the critical region.This technical difficulty is circumvented by using the multiple histogram reweighting proposed by(ρ,u)for a continuous Ferrenberg and Swensen26,27,28.With this method one can obtains Pβ,µL(ρ,u),i=1,...,R obtained set of values of(β,µ)from the knowledge of R histograms Pβi,µiLby performing R distinct MC simulations in the R(neighbor)thermodynamic states(βi,µi).Since the precision of the simulations offluid systems has still not reached that obtained in the MC simulations of Ising like systems it is impossible to construct ex nihilo thefixed point universal distribution P∗(x).In refs.29,30our attempts to match P L(M)on P∗(x) were realized by using the estimate of P∗is(x)made by Hilfer and Wilding32for the3D Ising model.Two new-and better-estimates of P∗is(x)obtained by Tsypin and Bl¨o te33for the 3D Ising model and the spin-1Blume-Capel model were considered in ref.31.The discussion is postponed to next section.IV.RESULTSA.General discussionIt turns out that thefield mixing parameter s of the RPM is practically independent of the temperature and of the size L of the system.Its magnitude,s∼−1.4629,30,31,is much higher than for neutralfluids(typically s∼0.02for square well or Lennard-Jonesfluids34) which explains the large dissymmetry of the liquid-vapor coexistence curve of the RPM.The collapse of the ordering operator distribution P L(M)onto the universal ordering distribution P∗is(x)given by the Blume-Capel model33is depicted in Fig.1for four different values of the volume ranging from V/σ3=5000to V/σ3=40000,i.e.up to a linear size L/σ=34.At volume V/σ3=5000a mismatch is observed at the lowest values of M due to an inadequate sampling of of the low density configurations at small volume.The overall good agreement leads us to conclude that the universality class of the RPM is that of the3D Ising model.The reduced apparent critical temperature T∗c(L)versus the size L of the system(in reduced units)has been plotted in Fig.2.Depending on the choice made for the universal ordering distribution P∗is(x)one obtains two sets of values of T∗c(L)from which T∗c(∞)can be obtained by making use of eq.(3.7).One obtains T∗c(∞)=0.04917±0.00002using P∗is(x)derived from the Blume-Capel model and T∗c(∞)=0.04916±0.00002using P∗is(x) obtained for the3D-Ising model.The approximate P∗is(x)of Hilfer and Wilding yields slightly different results.Note that in all cases we have used the Ising valuesν=0.63035 andθ=0.5336of the critical exponents.The previous analysis merely establishes the compatibility of the MC data with an Ising-like criticality.One can try to go beyond by considering the scaling behavior of the Binder cumulant<δM2>2LQ B(L)=•Conversely,fixing Q c=0.623andθ=0.53one obtainsβc=1/0.04918andν∼0.63±0.03.The variations of Q B(L)as a function ofβfor the different volumes is shown in Fig.3. Although there is considerable spread in the intersection points due to correction-to-scaling contributions,the corresponding values of Q c are close to the Ising value Q c=0.623and permit to rule out meanfield behavior(i.e.Q c=0.45738).Further support for Ising criticality is provided by the behavior of<δM2>at T∗c(L). According to the scaling hypothesis(3.6)it should scales as L2β/νwith system size.From the slope of the curve displayed in Fig.4one obtainsβ/ν=0.52in accord with the3D Ising value(0.517)and in clear contrast with the classical value1.In summary,our fss analysis leads to an estimate of the critical exponentsνandβ/νand the Binder cumulant Q c based on the sole knowledge of the critical temperature and the renormalization exponentθ.Within the numerical uncertainties these values are compatible with Ising-like criticality.Our conclusion is that the RPM,as ordinary neutralfluids,belongs to the universality class of the Ising model.A complete discussion of our MC data is out of the scope of the present paper and can be found in ref.31.For completeness I give below the values obtained for the critical temperature,chemical potentials and densities(the critical pressure is largely unknown):•T⋆c=0.04917±0.00002•ρ⋆c=0.080±0.005•µ⋆c=−13.600±0.005B.The specific heatThe revised scaling theory of Rehr and Mermin which is the framework of our fss analysis is however not the most general scaling theory which can be proposed for afluid system lacking the”particle-hole”symmetry.Its main weakness,as recognized already by Rehr and Mermin23,Yang and Yang39,and more recently by Fisher and co-workers40,41,42,is that it assumes the analyticity of the chemical potential at coexistenceµ(T)at the critical point. The more general scaling assumption should yield singularities for bothµ(T)and p(T)asT∗→T∗c.Let us examine the consequences of these singularities on the behavior of the specific heat capacity at constant volume C V.In the two phase region it can be rewritten as39C V=V T ∂2p∂T2V=C p+Cµ,(4.3) where C p(not to be confused with the specific heat capacity at constant pressure)and Cµ(not to be confused with the specific heat capacity at constant chemical potential)denote the two contributions to C V.I stress that,in eq.(4.3)p(T)andµ(T)denote the pressure and the chemical potential at coexistence.The formula can be used for any densityρg(T)<ρ<ρl(T)within the two phase region(ρg(T)andρl(T)being the densities of the gas and the liquid at coexistence respectively).In the revised scaling theory only C p diverges as |T∗−T∗c|−αwhereas one expects a divergence of both C p and Cµ(both as|T∗−T∗c|−α).In Fig.5I display the curves Cµ(T)and C V(T)along the locusχNNN∝<(N−<N>)3>=0,(4.4) for the four volumes considered in our last MC simulations31.Although the peak posi-tions shift correctly as∝L−1/νwith system size,in accord with fss theory18,19,there is no detectable scaling of the heights of the peaks which should scale as Lα/νwith L.These ob-servations corroborate similar results obtained by Valleau and Torrie43,44.In particular Cµdoes not show any anomaly which should challenge the use of eqs.(3.1)for the scalingfields.A possible explanation for the non singular behavior of C V(T)is that the amplitude of the singular term in C V(T)is small in the RPM and that the specific heat is dominated by its regular part.Note however that,the peak heights in C V(T)/V would scale,assuming Ising value forαonly by a factor2α/ν∼1.12when doubling the linear dimensions of the system. It is possible that such a small effect is not detectable within the statistical uncertainties of our calculations.V.CONCLUSIONIn this paper which resumes my talk at the Lviv NATO workshop I have described recent attempts to elucidate the nature of the critical behavior of the RPM model for ionicfluids,prototype of a system governed by long range Coulomb interactions by means of MC simulations.After endeavor over more than a decade we have now reached a point where we can claim confidently that the RPM belongs to the same universality class than the3D Ising model.The critical values of non-universal quantities such as the temperature and the chemical potential were established with a high accuracy whereas the uncertainties on the critical density are more significant,and the critical pressure is unknown.The behavior of the constant volume specific heat gives no indication of the expected Lα/νscaling of the peak height within the range of system sizes considered in the most recent simulations.Recent investigations of Camp and co-workers45where differences in the behavior of C V in the canonical and the GC ensemble are reported have emphasized this problem.At the moment it is difficult to explain this unexpected behavior of the specific heat.I have only discussed the properties of the continuous version of the RPM.The phase diagram of the various lattice versions of the model is in fact more complex46,47and was not described here due to a lack of place.I have also excluded from my presentation assymetric versions,either in charge or/and in size,of the continuum or lattice versions of the primitive model.The interested reader should consult recent works of Panagiotopoulos et al.48,49,50 and de Pablo et al.51,52,53.AcknowledgmentsI thank the organizers of the NATO workshop”Ionic Soft Matter”held in Lviv,Ukraine in April2004,Pr D.Henderson and Pr M.Holovko for having invited me to give a talk.It is a pleasure to acknowledge many scientific discussions with Pr.I.M.Mryglod,O.Patsahan, J.-P.Badiali,P.J.Camp,W.Schr¨o er,and J.Stafiej.Figure Captions•figure1:Collapse of the ordering distribution P L(M)onto the universal Ising order-ing distribution P∗is(x)for V/σ3=5000,T∗c(L)=0.004934,s=−1.45;V/σ3=10000, T∗c(L)=0.004926,s=−1.465;V/σ3=20000,T∗c(L)=0.004921,s=−1.47;and V/σ3=40000,T∗c(L)=0.004922,s=−1.43. P∗is(x)(solid circles)if the MC result of Tsypin and Bl¨o te(Ref.33)obtained for the Blume-Capel model.The scaling variable is x=a−1Lβ/νδM where a M is chosen in such a way that P L(x)has unit variance.M•figure2:The apparent critical temperature T∗c(L)as a function of L−(θ+1)/νwithθ=0.53,ν=0.630obtained by matching the universal ordering distribution calculatedfor the Blume-Capel model(top)and the Ising model(bottom).Extrapolating by linear least squarefit to the infinite volume limit yields T∗c(∞)=0.04917±0.00002 (top)and T∗c(∞)=0.04916±0.00002(bottom).•figure3:Variation of Q B(L)as a function of the inverse temperatureβfor the different volumes considered in ref.31.From top to bottom V/σ3=40000,20000,10000, and5000respectively.The symbols are the MC data and the lines thefits obtained by means of eq.(4.2).•figure4:Variations of ln<δM2>at T∗c(L)as a function of ln L.The slope of the linear least squarefit is2β/ν=1.04.•figure5:Variations of the total specific heat at constant volume C V/V and the contribution Cµ/V with temperature along the locusχNNN=0at volumes V/σ3= 5000,10000,20000,and40000(from left to right).∗Electronic address:Jean-Michel.Caillol@th.u-psud.fr1Weing¨a rtner H.,Schr¨o er W.//Adv.Chem.Phys.,2001,vol.116,p.1.2Fisher M.E.//J.Stat.Phys.,1994,vol.75,p.1.3Stell G.//J.Stat.Phys.,1995,vol.78,p.197.4Fisher M.E.//J.Phys.:Condens.Matter,1996,vol.8,p.9103.5Stell G.//J.Phys.:Condens.Matter,1996,vol.8,p.9329.6Vorontsov-Vel’Yaminov P.N.,Chasovkikh B.P.//High.Temp.,1975,vol.13,p.1071.7Chasovkikh B.P.,Vorontsov-Vel’Yaminov P.N.//High.Temp.,1976,vol.14,p.1174.8Valleau J.//J.Chem.Phys.,1991,vol.95,p.584.9Panagiotopoulos A.Z.//Fluid Phase Equilibria,1992,vol.76,p.97.10Panagiotopoulos A.Z.//Mol.Phys.,1987,vol.61,p.813.11Orkoulas G.,Panagiotopoulos A.Z.//J.Chem.Phys.,1994,vol.101,p.1452.12Caillol J.-M.//J.Chem.Phys.,1994,vol.100,p.2169.13Caillol J.-M.//J.Phys.:Condens.Matter,1994,vol.6,p.A171.14Caillol J.-M.,Weis J.-J.//J.Chem.Phys.,1995,vol.102,p.7610.15Brush S.G.,Sahlin H.L.,Teller E.//J.Chem.Phys.,1966,vol.45,p.2102.16de Leeuw S.W.,Perram S.W.,Smith E.R.//Proc.R.Soc.London A,1980,vol.373,p.27. 17Caillol J.-M.//J.Chem.Phys.,1999,vol.111,p.6528.18Privman V.,ed.,Finite Size Scaling and Numerical Simulation in Statistical Systems,Singapore, World Scientific,1990.19Cardy J.L.,ed.,Finite Size Scaling,Amsterdam,North Holland,1988.20Bruce A.D.,Wilding N.D.//Phys.Rev.Lett.,1992,vol.68,p.193.21Wilding N.D.,Bruce A.D.//J.Phys.:Condens.Matter,1992,vol.4,p.3087.22Wilding N.D.//Phys.Rev.E,1995,vol.52,p.602.23Rehr J.J.,Mermin N.D.//Phys.Rev.A,1973,vol.8,p.472.24Bruce A.D.//J.Phys.C:Solid State Phys.1981,vol.14,p.3667.25Nicolaides D.,Bruce A.D.//J.Phys.A:Math.Gen.1988,vol.21,p.233.26Ferrenberg A.M.,Swendsen R.R.//Phys.Rev.Lett.,1988,vol.61,p.2635.27Ferrenberg A.M.,Swendsen R.R.//Phys.Rev.Lett.,1989,vol.63,p.1195.28Deutsch H.-P.//J.Stat.Phys.,1992,vol.67,p.1039.29Caillol J.-M.,Levesque D.,Weis J.-J.//Phys.Rev.Lett.,1996,vol.77,p.4039.30Caillol J.-M.,Levesque D.,Weis J.-J.//J.Chem.Phys.,1997,vol.107,p.1565.31Caillol J.-M.,Levesque D.,Weis J.-J.//J.Chem.Phys.,2002,vol.116,p.10794.32Hilfer R.,Wilding N.D.//J.Phys.A,1995,vol.28,p.L281.33Tsypin M.M.,Bl¨o te H.W.J.//Phys.Rev.E,2000,vol.62,p.73.34Caillol J.-M.//J.Chem.Phys.,1998,vol.109,p.4885.35Ferrenberg A.M.,Landau.D.P.//Phys.Rev.B,1991,vol.44,p.5081.36Chen J.H.,Fisher M.E.,Nickel B.G.//Phys.Rev.Lett.,1982,vol.48,p.630.37Bl¨o te H.W.J.,Luijten E.,Heringa J.R.//J.Phys.A,1995,vol.28,p.6289.38Luijten E.,Bl¨o te H.W.J.//Phys.Rev.Lett.,1996,vol.76,p.1557.39Yang C.N.,Yang C.P.//Phys.Rev.Lett.,1964,vol.13,p.303.40Fisher M.E.,Orkoulas G.//Phys.Rev.Lett.,2000,vol.85,p.696.41Orkoulas G.,Fisher M.E.,Ust¨u n C.//J.Chem.Phys.,2000,vol.113,p.7530.42Orkoulas G.,Fisher M.E.,Panagiotopoulos A.Z.//Phys.Rev.E,2001,vol.63,p.051507.43Valleau J.,Torrie G.//J.Chem.Phys.,1998,vol.108,p.5169.44Valleau J.,Torrie G.//J.Chem.Phys.,2002,vol.117,p.3305.45Daub C.D.,Camp P.J.,Patey G.N.//J.Chem.Phys.,2003,vol.118,p.4164.46Luijten E.,Fisher M.E.,Panagiotopoulos A.Z.//J.Chem.Phys.,2001,vol.114,p.5468.47Panagiotopoulos A.Z.//J.Chem.Phys.,2002,vol.116,p.3007.48Panagiotopoulos A.Z.,Fisher M.E.//Phys.Rev.Lett.,2002,vol.88,p.045701-1.49Cheong D.W.,Panagiotopoulos A.Z.//J.Chem.Phys.,2003,vol.119,p.8526.50Romero-Enrique J.M.,Rull L.F.,Panagiotopoulos A.Z.//Phys.Rev.E,2003,vol.66,p.041204. 51Yan Q.,de Pablo J.J.//J.Chem.Phys.,2001,vol.114,p.177.52Yan Q.,de Pablo J.J.//Phys.Rev.Lett.,2002,vol.88,p.095504-1.53Yan Q.,de Pablo J.J.//J.Chem.Phys.,2002,vol.116,p.2697.00.00020.00040.00060.00080.049100.049200.049300.04940T*00.00020.00040.00060.0008L −(θ+1)/ν0.049100.049200.049300.04940T*。

PSA300 粒子尺寸分析仪操作说明书

Particle Size Distribution Analyzer TN168T T e e c c h h n n i i c c a a l l N N o o t t e e Powder DisperserEFFECT OF PSA300 POWDER DISPERSER ON SIZEAs in all particle analysis, sample selection and sample preparation remain critical to obtaining accurate results. This note addresses one aspect of sample preparation by showing the effects of varying disperser conditions on the obtainedparticle size distribution. In addition, we take advantage of the unique features of static image analysis to evaluate sample preparation strategies.IntroductionImage analysis is often considered areferee technique for particle sizing. The intuitive appeal of seeing particle picturesis compelling. In addition, the ability toextract more than size information, that is, shape, from images means that image analysis meets requirements not covered by other techniques. Naturally, methodsto improve image analysis results are important to extracting full value from this technique.The steps discussed here are examples ofthe steps used in developing a final method for analyzing a particular product. The HORIBA PSA300 Powder Disperser option is used to prepare microscope slides for static image analysis. With this device, sample particles are dispersedwith a controlled blast of air and allowedto settle on a microscope slide. In general,this method is quite gentle and distributesthe particles evenly across the microscope slide in a tightly controlled environment.It should be noted that not all static image analysis samples are best dispersed in this manner. For example, particle suspensions are often cast or spin coated onto the slide (1). Gels are spread onto a slide with a cover slip or razor blade (2). Some samples such as glass beads arefirst dispersed in a glycerin paste and then spread. But, for dry powders, the Powder Disperser is often the most convenient choice.In order to allow a single unit to be used with multiple sample types, the PSA300 Powder Disperser is quite flexible. In order to take advantage of this flexibility,the effect of various operating conditions on a particular sample type should beinvestigated. And, the effect of varying one operating condition, the starting pressure, is described here.Materials and MethodsA narrow size distribution fraction ofAvicel PH-101 microcrystalline cellulose was isolated by sieving. By using a narrow size fraction, one can ignore issues of sampling and counting a sufficient number of particles. The sample used here passed through 53 micron sieve openings, but not 45 micron sieve openings. For a discussion on reconciling sieving results with image analysis data (or dispensing with sieve analysis altogether), see (3).Two slides were prepared with the PowderDisperser. The only difference between the two slides was the starting pressure condition. The dispersion conditions are tabulated on the following page.Condition High PD Low PD Starting Pressure (torr)100 500 Dispersion FlowNormal Normal Dispersion Time (msec ) 500 500Air Restoration Delay (sec )30 30 Nozzles Medium, Large Medium,LargeThe starting pressure reflects the pressure of the chamber into which the particles are dispersed. The velocity of the blast of air can be controlled by lowering thechamber pressure since the air velocity is controlled by the difference between atmospheric pressure (760 torr) and chamber pressure. Of course, thedispersion flow setting is a different way to manipulate velocity. A more complete study would examine the effect of multiple disperser conditions. The slides aredesignated High PD and Low PD to reflect the pressure difference.A third slide, denoted “Manual,” wasprepared by using a spatula and dropping the powder onto the slide without applying any dispersion energy.The three slides were then examined with the HORIBA PSA 300 Static ImageAnalysis System using the 5x objective.ResultsQualitative AnalysisThe qualitative conclusions discussed here are based on reviewing images from each slide at a number of positions on the slide. The single photos presented here are for illustrative purposes even thoughqualitative analysis should be performedover multiple regions of a slide.A representative image from the High PD slide is shown below in Figure 1. From this image, one sees that the particles are not touching and therefore particle separation during image processing is unnecessary. Most notably for the discussion at hand, there are a substantial number of fine particles (less 20 microns) that are much smaller than the main particles.Figure 1 Representative image of particles dispersed under high pressure differenceconditions (high PD). Note the presence of a substantial number of fine particles that are an artifact of the dispersion process.Here one can take advantage of the nature of image analysis to confirm the presence of the fine particles in the original sample. An image from theManual slide is shown in Figure 2. In this image, the particles tend to overlapsignificantly; better dispersion will provide superior results to any numerical algorithm for particle separation.Therefore, this slide is not optimal for automated image analysis. Note the near absence of fine particles. It is clear that the fine particles observed in the High PD slide are not part of the original sample. This is consistent with the fact that the sample was prepared by sieving which would tend to remove any fine particles. From this image along with two or three others from the same slide, one candevelop a qualitative idea of the particlesize and shape for comparison withimages obtained from other slidesprepared with the HORIBA SampleDisperser.Figure 2 Representative image of manuallydispersed particles. Note the lack of fineparticles and the significant overlap of analyteparticles that preclude good automated imageanalysis.Finally, a typical image from the Low PDslide is shown in Figure 3. Note that in thiscase, unlike the High PD slide but similarto the Manual slide there are almost nofine particles. In addition, unlike theparticles in the Manual slide, the particlesare well separated; automated particleseparation during image processing isunnecessary. It should be pointed out thatthe number of particles in each imagecould be increased. And, optimizing thenumber of particles in each frame wouldbe the next step in method development.From these photographs, it is clear thatthe High PD dispersion conditions areaffecting the particles under study. Theparticles prepared under the Low PDdispersion conditions are unaffected bydispersion. Therefore, quantitative imageanalysis should be performed on theparticles prepared under the Low PDdispersion conditions.Figure 3 Representative image of particlesdispersed under low pressure differenceconditions (low PD). Note the lack of fineparticles and the distinct analyte particles.This sample is most appropriate for accurateautomated image analysis.Quantitative AnalysisLet us now compare the results of staticimage analysis from each sample. Weconsider two different distributionweightings: number weighted and volumeweighted. For the volume weighteddistribution, the particle volume isestimated based on the area of theparticle in the image and an assumedspherical form. Spherical volume is chosensince it is specified in ISO 13322-1 (4).The ellipsoidal volume calculation from thePSA 300 could be more appropriate.However, the choice of model does notaffect the conclusions drawn here.HighPDLowPDManualNumberMedianSize(microns)12 70 66VolumeMedianSize(microns)81 80 662Here it is clear that the number mediansizes (D50) obtained from the Manualslide, 66 microns, and the Low PD slide,70 microns are similar. For the High PD slide, the large number of fine particles brought the median size down to 12 microns. The small number of large agglomerates which are really overlapping or touching particles did not substantially change the obtained median size of the Manual slide. So, that value was still accurate.On the other hand, the volume median sizes (D50) obtained from the High PD slide, 81 microns and Low PD slide, 80 microns, were quite similar. This is because the volume fraction of fine particles is small. But the volume of large agglomerates was substantial enough to significantly perturb the measured volume median particle size from the Manual slide.The small difference between the number median and volume median particle sizes for the Low PD slide is a direct consequence of using a sample with a narrow size distribution.Since these numerical results were obtained by analyzing 392 images, and from 4000 to 20000 particles, thestatistics are certainly better than those from manually inspecting four or five images. For a discussion on the effect ofthe number of analyzed particles on the accuracy of the determined size distribution parameters such as the median size, see references (4) and (5). Review of only the numerical results does not clearly show which set of results is correct. But manual inspection of only afew images did clearly show which set of numbers is most accurate.ConclusionsImage analysis allows inspection of the results of different dispersion settings and this feature should be exploited in order to evaluate the quality of the slides prepared for image analysis. Comparing imagesfrom an undispersed sample to images from a dispersed sample rapidly verifies the appropriateness of dispersion conditions.References(1) AN193 Measuring 10 Micron PSL onthe PSA300, available from/us/particle(2) AN190 Particle Characterization of Ointments and Creams Using Image Analysis, available from/us/particle(3) AN142 Determination of the Roundness of Globules in the Pharmaceutical Industry, available from /us/particle(4) ISO 13322-1, Particle Size Analysis– Image Analysis Methods – Part 1: Static Image Analysis Methods(5) TN155 The Effect of Sample Size on Result Accuracy using Static Image Analysis, available from/us/particleCopyright 2011, HORIBA Instruments, Inc. For further information on this documentor our products, please contact:HORIBA Ltd.2, Miyanohigashi,KisshoinMinami-Ku Kyoto 601-8510 Japan+81 75 313 8121HORIBA Scientific34 BunsenIrvine, CA 92618 USA1-888-903-5001HORIBA Jobin Yvon S.A.S.16-18, rue du Canal - 91165 Longjumeau FranceTel. +33 (0)1 64 54 13 00/us/particle******************。

ceramic hydroxyapatite40um与80um区别 -回复

ceramic hydroxyapatite40um与80um区别-回复Ceramic hydroxyapatite is a biomaterial that mimics the natural composition of bones and teeth. It is commonly used in medical and dental applications due to its biocompatibility and osteoconductive properties. One of the factors that affect the performance of ceramic hydroxyapatite is its particle size. In this article, we will discuss the difference between ceramic hydroxyapatite with particle sizes of 40um and 80um.Particle size is a critical parameter in ceramic hydroxyapatite as it can influence factors such as surface area, porosity, and mechanical strength. A smaller particle size generally results in a larger surface area, which can enhance its interaction with biological tissues. On the other hand, a larger particle size can promote better mechanical stability. Let's delve deeper into the differences between ceramic hydroxyapatite with particle sizes of 40um and 80um.Firstly, the 40um ceramic hydroxyapatite has a significantly smaller particle size compared to the 80um variant. This means that the 40um hydroxyapatite particles have a larger surface area per unitweight, making it more reactive and potentially more advantageous for applications that require increased bioactivity, such as bone grafts or dental implants. The larger surface area can facilitate better adhesion and integration with the surrounding tissues, allowing for faster and more efficient healing.Secondly, the smaller particle size of 40um ceramic hydroxyapatite may also result in higher porosity. Porosity refers to the presence of open spaces or pores within the material, which can facilitate the infiltration of cells and blood vessels. A higher porosity can promote better nutrient exchange and tissue ingrowth, leading to improved integration and long-term stability of the implant. This characteristic makes the 40um variant more suitable for applications where rapid tissue regeneration and vascularization are desired.On the other hand, the 80um ceramic hydroxyapatite has a larger particle size, which imparts better mechanical stability and handling properties. Due to its larger particle size, the 80um variant exhibits a higher packing density and improved resistance to compression forces. This makes it more suitable for load-bearing applications, such as bone prosthetics or bioceramic coatings fordental implants.Additionally, the larger particle size of 80um ceramic hydroxyapatite may result in lower porosity compared to the 40um variant. While this reduces the material's ability to facilitate rapid tissue ingrowth, it may also hinder nutrient exchange and delayed the healing process. Therefore, the 80um variant is often used in applications where mechanical strength takes precedence over rapid tissue regeneration, such as orthopedic implants inweight-bearing bones.In conclusion, the choice between ceramic hydroxyapatite particle sizes of 40um and 80um depends on the specific requirements of the application. The 40um variant offers a larger surface area and higher porosity, making it ideal for situations that prioritize bioactivity and rapid tissue regeneration. On the other hand, the 80um variant provides better mechanical stability and is suitable for load-bearing applications. Understanding the differences between these two particle sizes allows for the selection of the most appropriate ceramic hydroxyapatite material for a particularmedical or dental need.。

FEM与DEM结合


1
1. Introduction
Fluid flow through particulate media is pivotal in many industrial processes, e.g. in fluidized beds, granular storage, industrial filtration and medical aerosols. Flow in these types of media is inherently complex and challenging to simulate, especially when the particulate phase is mobile. For the past two decades, particulate flows have been an active area of research and two widely used approaches are now considered state of the art. The first approach is based on an Eulerian continuum model of two phase flows, which only describes the averaged behavior of the multiphase media, see for example Kuipers et al. [1]. The second approach is based on an Eulerian-Lagrangian approach using finite volume/finite difference methods on a fixed grid as a fluid solver and either immersed boundary (IB) [2], fictitious domain (FD) [3], marker and cell (MAC) [4] or discrete element method (DEM) [5] for the particles. Both one-way and two-way couplings have been explored using these methods. While many fluid solvers are based on a staggered grid finite difference method, others e.g. Ladd [6, 7], Han et al. [8] and Feng et al. [2] have successfully utilized the lattice Boltzmann method (LBM) as a fluid solver for particle-fluid suspensions. The LBM is an attractive alternative due to its ease of implementation and parallelization; however, the selection of the lattice resolution, the collision kernel and accurate numerical implementation of various boundary conditions are still challenging [9]. A detailed description of flow through particulate media and accurate particle tracking can be obtained using discrete particle modeling (DPM) as proposed by Tsuji et al. [10], Kuipers et al. [11], Xu et al. [5] and Wu et al. [12]. In DPM, individual particles are tracked using Newton's laws of motion and particle-particle/wall interactions are taken into account on the (smallest) scale of the contacts. These models invariably couple a continuum solver for fluid with DEM, as originally proposed by Cundall & Strack [13], for particles. The coupling between fluid and particles is explicit and is achieved using semi-empirical drag laws or closure relations of fluid-particle interactions, e.g. Ergun et al. [14], Gidaspow [15], Drummond et al. [16], Gebart et al. [17]. In a recent studyfor two-way fluid-particle coupling on an unstructured mesoscopically coarse mesh is presented. In this approach, we combine a (higher order) finite element method (FEM) on the moving mesh for the fluid with a soft sphere discrete element method (DEM) for the particles. The novel feature of the proposed scheme is that the FEM mesh is a dynamic Delaunay triangulation based on the positions of the moving particles. Thus, the mesh can be multipurpose: it provides (i) a framework for the discretization of the Navier-Stokes equations, (ii) a simple tool for detecting contacts between moving particles, (iii) a basis for coarse graining or up-scaling and (iv) coupling with other physical fields (viz. temperature, electromagnetic, etc.). This approach is suitable for a wide range of dilute and dense particulate flows, since the mesh resolution adapts with particle density in a given region. Two-way momentum exchange is implemented using semi-empirical drag laws akin to other popular approaches, e.g. the discrete particle method, where a finite volume solver on a coarser, fixed grid is utilized. We validate the methodology with several basic test cases, including single- and doubleparticle settling with analytical and empirical expectations, and flow through ordered and random porous media, as compared against finely resolved FEM simulations of flow through fixed arrays of particles.

颗粒团聚分布的复合材料数值模型

第16卷第4期精密成形工程2024年4月JOURNAL OF NETSHAPE FORMING ENGINEERING79颗粒团聚分布的复合材料数值模型翁琳1*,吴爱中1,张镇国2,吴军梅1(1.上海工程技术大学,城市轨道交通学院,上海 201620;2.西安航天动力技术研究所固体推进全国重点实验室,西安 710025)摘要:目的提出颗粒团聚分布的数值模型参数化建模方法,以高效建立颗粒有序分布的复合材料微观构型。

方法以颗粒尺寸、颗粒团聚区域的半径以及团聚区域局部体积分数和整体体积分数的比值作为表征颗粒团聚结构的基本建模控制参数,首先建立了圆形/球形区域内的V oronoi多胞元结构,其次编写基于Python语言的接口程序,根据角点坐标和几何拓扑关系将多胞元结构导入有限元软件中,并对每个胞元按团聚区域局部体积分数进行缩放,形成了圆形/球形区域的颗粒团簇,最后使用随机顺序吸附算法将团聚颗粒在代表单元内周期性随机投放,生成了颗粒团聚分布的复合材料二维/三维微观构型。

结果使用不同控制参数进行数值建模测试,生成一个颗粒数目在1 000以内的二维/三维颗粒团聚微观构型仅需要数分钟,验证了本文方法的有效性。

使用二维模型对单轴拉伸进行数值模拟,结果表明,材料宏观塑性随着颗粒团聚程度的增大而降低。

结论本文建模方法可以为复合材料结构设计提供有效的数值仿真支撑。

关键词:复合材料;颗粒团聚;V oronoi结构;有限元方法;力学性能DOI:10.3969/j.issn.1674-6457.2024.04.010中图分类号:TB332 文献标志码:A 文章编号:1674-6457(2024)04-0079-08Numerical Model of Composite Materials with Agglomerated ParticlesWENG Lin1*, WU Aizhong1, ZHANG Zhenguo2, WU Junmei1(1. School of Urban Railway Transportation, Shanghai University of Engineering Science, Shanghai201620, China; 2. National Key Laboratory of Solid Rocket Propulsion, Institute of Xi'an AerospaceSolid Propulsion Technology, Xi'an 710025, China)ABSTRACT: The work aims to propose a parameterized modeling method for numerical models with agglomerated particles, in order to improve the efficiency of modeling the microstructure of composite material with ordered particle arrangement. The particle size, radius of particle aggregation region, and ratio of local volume fraction to overall volume fraction were taken as the basic modeling control parameters for characterizing particle aggregation structure. Firstly, a V oronoi multi cell tessellation was established in a circular/spherical region. Then, an interface program based on Python language was written to import the multi cell structure into finite element software based on corner coordinates and geometric topology relationships. Meanwhile, each cell was scaled according to the local volume fraction of the agglomeration area to form circular/spherical particle clusters. Fi-nally, a random order adsorption algorithm was used to place the agglomerated particles periodically in the representative units.收稿日期:2024-02-27Received:2024-02-27基金项目:国家自然科学青年基金(51701118)Fund:The National Natural Science Foundation of China (51701118)引文格式:翁琳, 吴爱中, 张镇国, 等. 颗粒团聚分布的复合材料数值模型[J]. 精密成形工程, 2024, 16(4): 79-86.WENG Lin, WU Aizhong, ZHANG Zhenguo, et al. Numerical Model of Composite Materials with Agglomerated Particles[J]. Journal of Netshape Forming Engineering, 2024, 16(4): 79-86.*通信作者(Corresponding author)80精密成形工程 2024年4月Therefore, two-dimensional/three-dimensional micro configurations of particle reinforced composite materials with different agglomeration characteristics were generated. When the number of particles was less than a thousand, it took only a few minutes to generate a two-dimensional/three-dimensional model, which verified the effectiveness of the method proposed in this study.Numerical simulation of uniaxial stretching was carried out with two-dimensional model, and the results showed that the model proposed could analyze the phenomenon that the macroscopic plasticity of the material decreased with the increase in the degree of particle agglomeration. The modelling method proposed in this study can provide effective numerical simulation support for the design of composite material structures.KEY WORDS: composite materials; particle agglomeration; V oronoi tessellation; finite element method; mechanical property复合材料将2种或者2种以上的组分材料结合在一起,充分利用各组分材料的优势进行互补,产生协同效应,从而获得优异的综合性能,其中非连续增强金属基体复合材料除了具有强度高、弹性模量高、耐磨损和导电导热好等优异性能外,还具有宏观各向同性、容易二次加工且制造工艺简单、造价低廉等优点,具有巨大的应用潜力[1-4]。

  1. 1、下载文档前请自行甄别文档内容的完整性,平台不提供额外的编辑、内容补充、找答案等附加服务。
  2. 2、"仅部分预览"的文档,不可在线预览部分如存在完整性等问题,可反馈申请退款(可完整预览的文档不适用该条件!)。
  3. 3、如文档侵犯您的权益,请联系客服反馈,我们会尽快为您处理(人工客服工作时间:9:00-18:30)。

a r X i v :c o n d -m a t /0506760v 1 [c o n d -m a t .s t a t -m e c h ] 29 J u n 2005Finite-size scaling and particle size cutoffeffects in phase separating polydispersefluidsNigel B.WildingDepartment of Physics,University of Bath,Bath BA27AY,United Kingdom.Peter Sollich and Moreno FasoloKing’s College London,Department of Mathematics,Strand,London WC2R 2LS,United Kingdom.(Dated:February 2,2008)We study the liquid-vapor phase behaviour of a polydisperse fluid using grand canonical simula-tions and moment free energy calculations.The strongly nonlinear variation of the fractional volume of liquid across the coexistence region prevents naive extrapolation to detect the cloud point.We describe a finite-size scaling method which nevertheless permits accurate determination of cloud points and spinodals from simulations of a single system size.By varying a particle size cutoffwe find that the cloud point density is highly sensitive to the presence of rare large particles;this could affect the reproducibility of experimentally measured phase behavior in colloids and polymers.PACS numbers:64.70Fx,68.35.RhMany complex fluids such as colloidal dispersions,liq-uid crystals and polymer solutions are inherently polydis-perse in character:their constituent particles have an es-sentially continuous range of size,shape or charge.Poly-dispersity is of significant practical importance because it can affect material properties in applications ranging from coating technologies and foodstuffs to polymer pro-cessing [1].However,our understanding of the fundamen-tal properties of polydisperse fluids remains very limited compared with what we know about their monodisperse counterparts.The challenge arises because a polydis-perse fluid is effectively a mixture of an infinite number of particle belling each by the value of its poly-disperse attribute σ,the state of the system (or any of its phases)has to be described by a density distribution ρ(σ),with ρ(σ)dσthe number density of particles in the range σ...σ+dσ.The most common experimental situ-ation is that in which the form of the overall or “parent”distribution ρ0(σ)is fixed by the synthesis of the fluid,and only its scale can vary depending on the proportion of the sample volume occupied by solvent.One can then write ρ0(σ)=n 0f 0(σ)where f 0(σ)is the normalized parent shape function and n 0=N/V the overall parti-cle number density.Varying n 0at a given temperature corresponds to scanning a “dilution line”of the system.A central issue in the physics of polydisperse fluids is the nature of their phase behaviour:in order to process a polydisperse fluid one needs to know under which condi-tions it will demix and what phases will result.However,the phase behaviour of polydisperse systems can be con-siderably richer than that of monodisperse systems,due to the occurrence of fractionation [2–4]:at coexistence,particles of each σmay partition themselves unevenly between two or more “daughter”phases as long as–due to particle conservation–the overall density distribution ρ0(σ)of the parent phase is maintained.As a conse-quence,the conventional fluid-fluid binodal of a monodis-perse system splits into a cloud curve marking the onset of coexistence,and a shadow curve giving the density of the incipient phase;the critical point appears at the in-tersection of these curves rather than at the maximum of either [5].In this letter we describe a joint simulation and theo-retical study of a model polydisperse Lennard-Jones fluid in which the size of the particles influences not only the length-scale but also the strength ǫij of the interparticle potentials,as defined in (2)below.For the case of size-independent interaction strengths,the critical point lies very close to the maximum of the cloud curve [6],whereas for the present model we find that it is substantially be-low (see Fig.3),as is observed in many experiments on complex fluids (see e.g.[7])and simplified theoretical cal-culations [8].At the critical temperature,T c ,there then exists a finite density range where phase separation oc-curs on the dilution line.Most results shown below will be at this temperature;note that we will be interested mainly in the low-density part of the coexistence region rather than the critical effects at the other end,using T c merely as a convenient temperature scale.The simulations were performed within the grand canonical ensemble (GCE).This is particularly useful for polydisperse systems,where it permits sampling of many different realizations of the particle size distribu-tion while catering naturally for fractionation effects.Operationally,we ensure that the ensemble averaged den-sity distribution always equals the desired parent distri-bution ρ0(σ)by controlling an imposed chemical poten-tial distribution µ(σ).A combination of novel and exist-ing techniques [9]are required to tune µ(σ)such as to track the dilution line,i.e.to vary the parent density n 0but not its shape f 0(σ).In the GCE simulations,the number density n is a fluc-tuating variable with average equal to n 0.Its distribution p (n ),shown in Fig.1for a range of parent densities n 0at T =T c ,is a key observable.In the coexistence region it has two distinct peaks,which we sample using multi-liquid peak disappears,at exponentially small values of v l visible only on a log scale(Fig.1(b)).The observed variation of p(n)raises the question of how to detect the cloud point n0cl,defined as the lowest parent density n0where stable phase coexistence occurs. In a monodisperse system this is straightforward because the cloud point also gives the density of the gas phase, which remains constant throughout the coexistence re-gion.One then simply detects the point where the gas and liquid peaks have the same weight,i.e.r=v l/v g=1, and measures the gas density there.(The criterion r=1 has the added advantage of leading to only exponentially smallfinite-size corrections to the value ofµat coexis-tence[11]).However,this method fails in a polydisperse system because fractionation causes the densities and size distributions of the coexisting phases to vary with n0[5]. One could attempt to locate the cloud point instead by extrapolating in n0to the point where v l→0[6].But in our system the dependence of v l on n0is so strongly nonlinear—another effect of fractionation,see inset of Fig.1—that the resulting cloud point estimates would have very large error bars.Indeed,on a linear plot of v l FIG.2:Ratio r of liquid to gas fractional volumes on approach to the cloud point at T=T c forσc=1.4.The inset shows the(negative,scaled)second derivative of ln r w.r.t.n0.The peak position gives an estimate of the cloud point density. Squares indicate the scaled master curve(1).Forfinite L,metastable coexistence can still be ob-served in the density region n0<n0cl where∆P=O(1), but here r will be exponentially small.Fig.2showsthis effect clearly:r is independent of L for sufficientlylarge n0,but the curves depart rapidly from each other (note the log-scale)for smaller n0.The cloud point sep-arates the two regimes,permitting the estimate n0cl≈0.0825±0.0005for the parameters shown in thefigure.Toderive a method which can estimate n0cl even from datafor only a single system size L,we use the fact that∆P is O(1)and scales linearly with n0−n0cl to leading order near the cloud point,and hence ln r∼L d(n0−n0cl).This applies for n0<n0cl,while above n0cl one has ln r=O(1). Thus the derivative(∂/∂n0)ln r should drop from an O(L d)plateau to O(1)around n0=n0cl.In the second derivative−(∂/∂n0)2ln r this drop will manifest itself as a peak.A smooth derivative can be extracted from sim-3ulation data using histogram reweighting,and the peak position then serves as an estimate for n0cl.This is shownin the inset of Fig.2,and gives n0cl≈0.0823from the largest L,consistent with our earlier estimate derivedfrom comparing data for different L.The above arguments can be formalized using the re-sults of[11],which pertain to the monodisperse case but which we have generalized to polydisperse systems[12]. Wefind that for large L the second-derivative plot ap-proaches a universal master curve−∂(1+z)3,˜n0=z+ln z,(1)parameterized by z.The scaled parent density is defined here as˜n0=aL d(n0−n0cl)+ln(bL d n0cl)with a and b system-specific dimensionless scale factors.This scaling implies that the cloud point estimate from the peak posi-tion hasfinite-size corrections of order L−d ln L,while the peak width and height scale as L−d and L2d,respectively. Our data are consistent with the width and height scal-ing and with the dominant L−d dependence of the peak shifts[12].The master curve(1),appropriately scaled, is overlaid onto the largest-L data in Fig.2(inset)and shows excellent agreement.Fig.1shows that the metastable liquid peak in p(n) persists until well below the cloud point n0cl.The point at which it disappears marks the so-called meanfield spinodal[13]where the liquid is unstable to small den-sityfluctuations.The parent density n0sp where this oc-curs should tend to an L-independent value as L grows large[13],and our data(not shown)are consistent with this.Spinodals in monodisperse systems are convention-ally characterized by the density of the phase that be-comes unstable,which is located inside the region where phase separation occurs.Here we use instead the den-sity n0sp of the coexisting stable phase,which is outside this region.This is a more meaningful representation in the polydisperse context since only the stable(majority) phase has the parental size distribution,while that of the metastable(minority)phase is determined indirectly via chemical potential equality.Equipped with a systematic method for determining cloud points,we now consider the overall phase diagram of our system,the interparticle potential of which was assigned the Lennard-Jones form:u ij=ǫij (σij/r ij)12−(σij/r ij)6 (2) withǫij=σiσj,σij=(σi+σj)/2and r ij=|r i−r j|. The potential was truncated for r ij>2.5σij and no tail corrections were applied.The diametersσare drawn from a(parental)Schulz distribution f0(σ)∝σz exp[(z+1)σ/¯σ],with a mean diameter¯σwhich sets our unit length scale.We chose z=50,corresponding to a moderate degree of polydispersity:the standard de-viation of particle sizes isδ≡1/√4distributions at T=T c forσc=1.4,1.6,1.8,together with the parent distribution f0(σ).Inset:MFE theory prediction for larger cutoffσc=3;note the second peak in the shadow distribution.despite particle sizes aroundσc being very rare in the parent,they occur in significant concentration in the shadow liquid.Physically,since large particles interact more strongly,their presence leads to a substantial free energy gain at the shorter interparticle separations of the liquid.One is led to enquire whether the gas phase cloud point density would eventually tend to a nonzero limit asσc is increased.The inset of Fig.3shows the simulation results and theoretical predictions.The former exhibit a further strong decrease of n0cl by≈25%betweenσc= 1.6and1.8;the latter suggest that this trend continues and that the cloud point density tends to zero for large σc.Such an unusual effect has previously been seen in theoretical studies of polydisperse hard rods with wide length distributions[16]and is also predicted to occur in solid-solid phase separation of polydisperse hard spheres [17],though only for largeσc and distributions with fatter than exponential tails.Here the decrease of n0cl is clear even forσc of O(1),i.e.of the same order as¯σ,and scaling estimates suggest that cutoffeffects occur for any size distribution with tails heavier than a Gaussian[12]. The physical origin of the decrease of n0cl to zero is the appearance(for largeσc)of a second peak in the shadow phase size distribution nearσc(Fig.4,inset). As with the hard rods,we expect this second peak to eventually dominate asσc increases so that the shadow phase consists of ever more strongly interacting particles whose sizes are concentrated nearσc.We speculate that as a consequence,there exists some cutofffor which the shadow phase liquid freezes into a quasi-monodisperse crystal phase.Indeed our simulations provide evidence for this scenario:for the large cutoffσc=2.8the liquid spontaneously freezes to an f.c.c.crystal structure[12]. Although we observe this only for small n0values close to the spinodal point it is conceivable that,forσc values larger than those presently accessible to simulations,the freezing might occur from the stable liquid phase. Finally,with regard to the cloud curves as a whole (Fig.3),we note that significant cutoff-dependent shifts occur only for densities below the critical density.This is consistent with our interpretation above:for higher den-sities,the shadow phase is a gas of lower density than the parent.In this,the concentration of large particles is suppressed and that of small particles negligibly en-hanced because of their weak interactions.The shadow size distributions are therefore concentrated well within the range0.5...σc(data not shown)so that no cutoffdependence arises.In summary,the task of accurately locating cloud points of polydispersefluids via simulation is severely complicated by fractionation effects.We have presented a generally applicablefinite-size scaling method which addresses this problem.Application to a model poly-dispersefluid reveals the cloud curve to be highly sen-sitive to the presence of very rare large particles.Such effects imply that in experiments on polydispersefluids (see e.g.[4])it may be important to monitor and con-trol carefully the tails of the size(or charge,etc)distri-bution.Otherwise undetected differences could lead to large sample-to-samplefluctuations in the observed phase behaviour.[1]G.H.Fredrickson,Nature(London)395,323(1998).[2]R.M.L.Evans,D.J.Fairhurst,and W.C.K.Poon,Phys.Rev.Lett.81,1326(1998).[3]K.Ghosh and M.Muthukumar,Phys.Rev.Lett.91,158303(2003).[4]B.H.Ern´e et al,Langmuir21,1802(2005).[5]P.Sollich,J.Phys.Condens.Matter14,R79(2002).[6]N.B.Wilding.M.Fasolo and P.Sollich,J.Chem.Phys.121,6887(2004).[7]R.Kita,K.Kubota and T.Dobashi,Phys.Rev.E56,3213(1997).[8]L.Bellier-Castella,H.Xu and M.Baus,Phys.Rev.E113,8337(2000).[9]N.B.Wilding,J.Chem.Phys.119,12163(2003);N.B.Wilding and P.Sollich,J.Chem.Phys.116,7116(2002).[10]B. A.Berg and T.Neuhaus,Phys.Rev.Lett.68,9(1992).[11]C.Borgs and W.Janke,Phys.Rev.Lett.68,1738(1992);C.Borgs and R.Kotecky,Phys.Rev.Lett.68,1734(1992).[12]N.B.Wilding.M.Fasolo and P.Sollich,unpublished.[13]P.A.Rikvold et al,Phys.Rev.E49,5080(1994).[14]L.Kvitek et al,J.Mater.Chem15,1099(2005).[15]P.Sollich,P.B.Warren and M.E.Cates,Adv.Chem.Phys.116,265(2001).[16]A Speranza and P Sollich,J.Chem.Phys.118,5213(2003).[17]M.Fasolo and P.Sollich,Phys.Rev.E70,041410,(2004).。

相关文档
最新文档