• Nie Znaleziono Wyników

Determining the nucleation rate from the dimer growth probability

N/A
N/A
Protected

Academic year: 2021

Share "Determining the nucleation rate from the dimer growth probability"

Copied!
5
0
0

Pełen tekst

(1)

Determining the nucleation rate from the dimer growth probability

Joop H. ter Horsta兲

Laboratory for Process Equipment, Delft University of Technology, Leeghwaterstraat 44, 2628 CA Delft, The Netherlands

Dimo Kashchiev

Institute of Physical Chemistry, Bulgarian Academy of Sciences, ul. Acad. G. Bonchev 11, Sofia 1113, Bulgaria

共Received 10 June 2005; accepted 27 July 2005; published online 22 September 2005兲

A new method is proposed for the determination of the stationary one-component nucleation rate J with the help of data for the growth probability P2 of a dimer which is the smallest cluster of the nucleating phase. The method is based on an exact formula relating J and P2, and is readily applicable to computer simulations of nucleation. Using the method, the dependence of J on the supersaturation s is determined by kinetic Monte Carlo simulations of two-dimensional 共2D兲 nucleation of monolayers on the共100兲 face of Kossel crystal. The change of J over nearly 11 orders of magnitude is followed and it is found that the classical nucleation theory overestimates the simulation J values by an s-dependent factor. The 2D nucleus size evaluated via the nucleation theorem is described satisfactorily by the classical Gibbs-Thomson equation and its corrected version accounting for the spinodal limit of 2D nucleation. © 2005 American Institute of Physics. 关DOI:10.1063/1.2039076兴

I. INTRODUCTION

The determination of the stationary nucleation rate J is a central problem in nucleation theory and experiment.1,2This quantity gives the number of new-phase clusters with large enough size that are steadily nucleated per unit time and unit volume共or area兲 of the supersaturated old phase. Experimen-tally, the determination of J is not easy because of the diffi-culties in the realization of time-independent supersaturation ⌬␮ and of an old phase with well-defined properties 共e.g., purity, number and activity of nucleation-active centers, etc.兲. Such difficulties do not exist in computer simulations of nucleation and for that reason these are valuable tools for deeper understanding of the nucleation process at the mo-lecular level and for reliable verification of theoretical ex-pressions for the nucleation rate.

However, the direct determination of J by computer simulations is also accompanied with difficulties. A major problem is the excessively long computer time required by the presently available methods to obtain J when the nucle-ating system is subjected to the relatively low supersatura-tions used in real experiments. As the computer time for finding J shortens strongly with increasing⌬␮, these meth-ods are only applicable to systems at high enough supersatu-rations in a rather narrow range. There exists thus a necessity of new methods for direct determination of J, which would allow the simulations to be carried out within a reasonably short computer time, but for supersaturations in an extended range comprising relatively low⌬␮values.

The objective of the present paper is to propose a new method for direct determination of the stationary nucleation

rate by computer simulations of the growth probability of a dimer which is the smallest cluster that can decay to a single molecule 共or atom兲, i.e., to a monomer. In some cases this method may allow obtaining the J共⌬␮兲 dependence for such low supersaturations that are comparable with those used in real experiments.

II. THEORY

Nucleation is known to occur by the formation and growth of n-sized clusters of the new phase共n is the number of molecules constituting the cluster兲. According to the Sz-ilard model of nucleation, the clusters change their size by successive attachment and detachment of monomers to and from them. Within this model, the stationary nucleation rate

J 共m−3s−1or m−2s−1兲 is given by the Becker-Döring formula1–3

J = f1C1/共1 + S2兲, 共1兲

which is exact and applies to any kind of one-component nucleation 关homogeneous, heterogeneous, three-dimensional 共3D兲, two-dimensional 共2D兲, etc.兴. In Eq. 共1兲 f1共s−1兲 is the frequency of monomer attachment to a monomer, and

C1共m−3or m−2兲 is the concentration of monomers. For in-stance, in one-component homogeneous nucleation of drop-lets or crystals in a vapor C1共m−3兲 is the actual concentration of molecules in the vapor, and in 2D nucleation of liquid or crystalline monolayers on a solid substrate without nucleation-active centers on it C1共m−2兲 is the actual concen-tration of adsorbed molecules on the substrate. The quantity

S2is defined by

a兲Author to whom correspondence should be addressed. Electronic mail:

J.H.terHorst@wbmt.tudelft.nl

(2)

S2=

n=2 M−1

g2g3. . . gn/f2f3. . . fn, 共2兲

where fn共s−1兲 and gn共s−1兲 are, respectively, the frequencies

of monomer attachment to and detachment from an n-sized cluster, and M is the total number of molecules in the old phase.

On the other hand, the attachment and detachment fre-quencies govern the growth of an n-sized cluster to a mac-roscopically large cluster and the decay of the cluster to a monomer. Within the Szilard model, the corresponding growth and decay probabilities can be calculated4,5with the help of a result from probability theory.6 The growth prob-ability Pn is given by the expression

5共2艋n艋M −1兲

Pn= 1 − Sn/共1 + S2兲, 共3兲

which is exact and valid for any kind of one-component nucleation and in which

Sn=

i=n M−1

g2g3. . . gi/f2f3. . . fi. 共4兲

While Pnis not connected simply with J, the probability

P2of growth of a dimer to a cluster with macroscopic size is. Indeed, at n = 2 Eq.共3兲 yields

P2= 1/共1 + S2兲. 共5兲

Combining Eqs.共1兲 and 共5兲, we arrive at the following gen-eral relation between the stationary nucleation rate and the dimer growth probability:

J = f1C1P2. 共6兲

Within the Szilard model, this equation is also exact and applicable to any kind of one-component nucleation. It is a particular case of a more general equation found by White4 to relate J and Pn.

Equation共6兲 has a very simple physical meaning. It says that J is merely the product of the frequency f1C1with which monomers become dimers and the probability P2 that these dimers grow to microscopically large clusters rather than de-cay to monomers. This equation is very convenient for use in computer simulations, because then the monomer-to-monomer attachment frequency f1and the monomer concen-tration C1 are either known independently or obtainable by separate simulations. Thus, the only quantity that has actu-ally to be found by simulations is the dimer growth probabil-ity P2. This can be done for a rather wide supersaturation range including ⌬␮ values that in some cases may be as small as those used in real experiments. Moreover, employ-ing P2 for determination of J with the help of Eq. 共6兲 is advantageous, since the value of P2obtained by simulations is invariant with respect to the definition chosen for the clus-ters of size n = 3 , 4 , . . .. This is so, because the simulations follow the evolution of a single dimer to a large enough cluster capable of irreversible growth irrespectively of how its size is defined.

When the dependence of P2 on⌬␮ is obtained at con-stant temperature, it can be used for a model-independent determination of the nucleus size共the nucleus is that

particu-lar cluster that requires maximum work for its formation兲. This can be done with the aid of the nucleation theorem.2,7–11 Indeed, representing J in its general form1,2,12

J = A exp共− W*/kT兲, 共7兲

from Eq. 共6兲 we get

P2=共A/f1C1兲exp共− W*/kT兲, 共8兲

where A共m−3s−1or m−2s−1兲 is a kinetic factor, W* is the work to form the nucleus, T is the absolute temperature, and

k is the Boltzmann constant. Taking the logarithm of both

sides of Eq. 共8兲 and employing the nucleation theorem 共for one-component nucleation at constant T兲 in the form2,7,8,11

dW*/d

old= −⌬n* 共9兲

results in

⌬n*= kTd共ln P

2兲/d␮old− kTd关ln共A/f1C1兲兴/d␮old, 共10兲 where ␮old is the chemical potential of the old phase, and ⌬n*, the excess number of molecules in the nucleus, is the difference between the number of molecules in the spatial region occupied by the nucleus and the number of molecules in the same region before the nucleus formed. When ␮oldis changed at constant T, the quantities A, f1, and C1are given by2 A = A0exp共␮old/ kT兲, f1= f1,0exp共␮old/ kT兲, and C1 = C1,0exp共␮old/ kT兲, where the factors A0, f1,0, and C1,0may depend only weakly on ␮old. Hence, to a good approxima-tion, Eq. 共10兲 can be used in the form

⌬n*= kTd共ln P

2兲/d␮old+ 1. 共11兲

For example, in the case of one-component nucleation of droplets or crystals in vapors behaving as ideal gas, we have1,2␮old=␮ref+ kT ln p, whererefis a reference chemical potential and p is the pressure of the vapors. Equation 共11兲 then becomes

⌬n*= d共ln P

2兲/d共ln p兲 + 1. 共12兲

III. SIMULATION RESULTS

The kinetic Monte Carlo simulation results presented be-low are for one-component 2D nucleation of monolayers on the molecularly smooth surface of the共100兲 face of a Kossel crystal. The simulation details to determine the growth prob-ability Pn of an n-sized 2D cluster of monolayer thickness

can be found elsewhere.13The attachment frequency ka共s−1兲

of a molecule to whatever site on the crystal surface is given by13–15

ka= kees, 共13兲

where the dimensionless supersaturation s is defined by

s =⌬␮/kT =共␮old−␮e兲/kT, 共14兲

and ke共s−1兲 and ␮e are the values of ka and ␮old at phase equilibrium共s=0兲. As only nearest-neighbor interactions are taken into account, the detachment frequency kd,j共s−1兲 of a

molecule from a site with j = 0 , 1 , 2 , 3 , 4 lateral nearest neighbors is of the form13–15

(3)

kd,j= kee2␻共2−j兲, 共15兲

where ␻=␾/ kT is half the dimensionless overall nearest-neighbor binding energy 共␾is half the value of this overall energy兲. The simulations were performed at a fixed␻= 2, so that the system studied is equivalent to the 2D Ising model with a nearest-neighbor coupling equal to kT. The supersatu-ration values used were in the range of 0.6艋s艋7.5. A simu-lation run at a given value of s started with a single dimer on the crystal face and ended either when the dimer decayed to a monomer共the run was then qualified as negative兲 or when the dimer grew to a cluster of sufficiently large size nmax共the run was then qualified as positive兲. The dimer growth prob-ability P2 was calculated from the ratio of the number of positive runs to the total 共positive and negative兲 number of runs. The total number of runs to obtain P2at a given super-saturation was between about 5⫻103 at the largest s value and 4⫻108 at the smallest s value. The maximum cluster size nmax for dimer growth was set equal to 200, because it was found that clusters of this size practically did not decay to monomers.

In order to determine the nucleation rate J with the help of Eq.共6兲 we need to know f1 and C1. In view of Eq.共13兲, the monomer-to-monomer attachment frequency f1 is given exactly by

f1= 4ka= 4kees, 共16兲

because the monomer has four nearest-neighbor attachment sites at which a dimer can be created. The concentration C1 of monomers on the crystal face can be evaluated from the approximate expression14

C1=共1/a0兲es−4␻, 共17兲

which follows from Eqs. 共13兲 and 共15兲, since according to adsorption theory C1 is approximately the product of the in-cidence rate ka/ a0 of molecules and the average lifetime 1 / kd,0of an adsorbed molecule 共a0 is the area occupied by such molecule on the crystal face兲. Combining Eqs. 共6兲, 共16兲, and共17兲 leads to the formula

J共s兲 = 4共ke/a0兲e2s−4␻P2共s兲 共18兲

in which only the P2共s兲 dependence has to be obtained by simulations.

The circles in Fig. 1 represent P2as a function of s in the

supersaturation range studied. The standard deviation of the

P2 values is within the size of the circles. The dashed line visualizes the value 1 / 2 of P2. As seen, P2⬎1/2 for s ⬎3.5, which means that at such high supersaturations the dimer has a higher chance to grow to a large supernucleus than to decay to a monomer. Since the condition P2= 1 / 2 determines approximately the size n*of the nucleus,5

Fig. 1 reveals that at s⬇3.5 the dimer is the nucleus. Accordingly, for greater s values the dimer is already a supernucleus. We note also that in the determination of the nucleation rate it is physically meaningful to use only the P2 data for s⬍4.3, because this is the value共at␻= 2兲 of the spinodal supersatu-ration ss which limits the 2D nucleation on the crystal face

关4.3 is the s value at the spinodal point of the mean-field adsorption isotherm of a monolayer on the 共100兲 face of Kossel crystal at ␻= 2; see, e.g., Ref. 16兴. Figure 1 shows that, as expected on the basis of general considerations,17 P2 has no singularity at or near the mean-field spinodal super-saturation ss= 4.3.

The circles in Fig. 2 represent the nucleation rate J ob-tained from Eq.共18兲 with the help of the P2共s兲 data from Fig. 1. As seen, the present method of the determination of J allows following the change of J with s over nearly 11 orders of magnitude. The standard way to directly determine J by counting the number of supernuclei appearing per unit time does not cover such a wide range of J values. This is evident in Fig. 2 where the squares are the J共s兲 data obtained by Weeks and Gilmer15in this way. Figure 2 shows also that the present method for the determination of J is in very good agreement with the standard one. We note as well that the lowest s values in Fig. 2 compare with those used by Bostanov et al.18 in studying experimentally the 2D nucle-ation rate in electrochemical growth of the 共100兲 face of a silver crystal共these authors experimented at s=0.62–0.80兲.

We can now verify whether the classical nucleation theory describes adequately the J共s兲 data in Fig. 2. According to this theory,2 J is given by Eq.共7兲 with

A = zfn*C0, 共19兲

where fn* is the attachment frequency of monomers to the

nucleus, C0= 1 / a0is the concentration of nucleation sites on the crystal face共because the face is free of nucleation-active

FIG. 1. Supersaturation dependence of the dimer growth probability: circles—simulation data; dashed line—the value 1 / 2 of the probability; solid curve—Eq.共29兲.

FIG. 2. Supersaturation dependence of the 2D nucleation rate: circles— simulation data obtained from Eq.共18兲 with the help of the P2共s兲 data in Fig.

1; squares—simulation data of Weeks and Gilmer;15dashed curve—Eq.共25兲 of the classical nucleation theory; solid curve—Eq.共27兲; dotted curve—Eq. 共27兲 with right-hand side multiplied by 1/40.

(4)

centers兲, and the Zeldovich factor z is of the form

z =共W*/4kTn*21/2. 共20兲

As this theory treats the 2D nucleus as a square prism with monomolecular thickness and size-independent specific edge energy␬共J/m兲, the nucleation work W*and the nucleus size n*are given by2

W*= 42kT/s, 共21兲

n*= 4␧2/s2, 共22兲

where

␧ =␬a01/2/kT 共23兲

is the nucleus dimensionless specific edge energy. Also, the square nucleus has 4n*1/2 attachment sites which provide nearest-neighbor binding at its periphery so that, in view of Eq.共13兲, fn*is of the form

fn*= 4n*1/2kees. 共24兲

Combining Eqs.共7兲, 共19兲–共22兲, and 共24兲 yields the classical formula

J =共2/␲1/2兲共ke/a0兲s1/2e

s

exp共− 4␧2/s兲 共25兲

in which the factor 2 /␲1/2corrects the factor1/2/ 2 used by Weeks and Gilmer.15 Because of the square shape of the 2D nucleus, in Eqs. 共21兲, 共22兲, and 共25兲 ␧ is related with␻ by the expression19

␧ =␻+ ln关共1 − e−␻兲/共1 + e−␻兲兴, 共26兲

which gives the specific edge energy of an infinitely long 共10兲 step on the 共100兲 face of the Kossel crystal. Thus, Eqs. 共21兲, 共22兲, and 共25兲 represent the s dependence of W*/ kT, n*, and the dimensionless 2D nucleation rate Ja0/ kewithout free

parameters.

Equation 共21兲 does not satisfy the thermodynamic re-quirement for the annulment of the nucleation work W*at the spinodal supersaturation. Recently, it was shown2,16,20–24that in various cases of nucleation this annulment can be ac-counted for by the introduction of a correction factor in the corresponding classical formula for W*. For the case of 2D nucleation considered here, the correction factor is of the form 共1−s2/ ss

2兲 共see Ref. 16兲 and appears in the right-hand side of Eq. 共21兲. Due to Eq. 共7兲, the nucleation rate is also affected by this correction so that Eq.共25兲 becomes16

J =共2/␲1/2兲共ke/a0兲s1/2esexp关− 共4␧2/s兲共1 − s2/ss2兲兴. 共27兲

The dashed curve in Fig. 2 displays the J共s兲 dependence from Eq. 共25兲 with ␧=1.73, the value following from Eq. 共26兲 at␻= 2. We observe that the classical nucleation theory overestimates by an s-dependent factor the nucleation rate determined by the simulations. The disagreement is greater than that found by Weeks and Gilmer,15because in Eq.共25兲 these authors used ␧=␻ which is the low-temperature ap-proximation for␧ following from Eq. 共26兲 for␻Ⰷ1 关physi-cally, this approximation corresponds to neglecting the mo-lecular roughness of the 共10兲 step at T⬎0兴. The corrected

J共s兲 dependence from Eq. 共27兲 with spinodal supersaturation ss= 4.3 is illustrated by the solid curve in Fig. 2. As seen,

quantitatively, the description of the simulation data by Eq. 共27兲 is not better than that by Eq. 共25兲. Qualitatively, how-ever, the description is improved, because J from Eq.共27兲 is greater than J from the simulations by a practically

s-independent factor. This improvement is visualized by the

dotted curve in Fig. 2, which exhibits the J values calculated from Eq. 共27兲 and multiplied by 1/40. Why does this factor improve the performance of Eq.共27兲 remains an open ques-tion.

We can now use Eq.共10兲 to determine the size of the 2D nucleus with the help of the P2共s兲 dependence obtained by the simulations. According to Eq.共25兲, for the kinetic factor

A we have共2/␲1/2兲共k

e/ a0兲s1/2esso that, in view of Eqs.共14兲, 共16兲, and 共17兲, for the simulation model Eq. 共10兲 takes the form

⌬n*= d共ln P

2兲/ds + 1 − 1/2s 共28兲

which gives⌬n* more accurately than Eq.共11兲. The deriva-tive in Eq.共28兲 can be evaluated with the aid of the function ln P2= − 10.2127/s − 1.6202 + 2.1843s − 0.3083s2 共29兲 providing the best fit to the simulation data in Fig. 1 and depicted by the curve in this figure. With the so-evaluated derivative, Eq. 共28兲 yields the ⌬n*共s兲 dependence repre-sented by the circles in Fig. 3. For comparison, the squares in Fig. 3 illustrate the⌬n*values obtained elsewhere2

from the

J共s兲 data of Weeks and Gilmer,15 and the triangles show the

n*共s兲 values determined by ter Horst and Jansens13

by simu-lations of the cluster growth probability at ␻= 2. As seen, despite that the number n* of molecules in the 2D nucleus depends on the cluster definition used by the latter authors, it is close to the definition-independent excess number⌬n* of molecules in the nucleus. Also, the nucleus size at the lower supersaturations in Fig. 3 is comparable with that operative in real experiments: n*= 19– 32 was found2to correspond to the s range used in the experiments of Bostanov et al.18on 2D nucleation in electrochemical growth of the共100兲 face of a silver crystal.

Recently, it was shown16that for 2D nucleation in crys-tal growth the⌬n*共s兲 dependence can be expressed approxi-mately as

FIG. 3. Supersaturation dependence of the nucleus size: circles—simulation ⌬n*共s兲 data obtained from Eq. 共28兲; squares—⌬n*共s兲 data from Ref. 2,

which follow from the simulation J共s兲 dependence of Weeks and Gilmer;15 triangles—n*共s兲 data of ter Horst and Jansens13obtained by simulations of

the cluster growth probability; dashed curve—Gibbs-Thomson Eq. 共22兲; solid curve—Eq.共30兲.

(5)

⌬n*= n*共s兲共1 + s2/s

s

2兲, 共30兲

where n*共s兲 is given by the Gibbs-Thomson Eq. 共22兲. The solid curve in Fig. 3 exhibits the ⌬n*共s兲 dependence pre-dicted by Eq. 共30兲 with n* from Eq. 共22兲 and ss= 4.3. For

comparison, the dashed curve displays the nucleus size n* from Eq.共22兲. As seen, Eq. 共30兲 provides a better description of the simulation data than the Gibbs-Thomson Eq. 共22兲. Except for s = 4.25, the⌬n* values from Eq.共30兲 are within 17% of those from the simulations.

IV. CONCLUSION

In the scope of the Szilard model, the stationary rate J of one-component nucleation has a simple probabilistic inter-pretation: according to the exact Eq. 共6兲, for any kind of nucleation J is the product of the frequency f1C1with which monomers become dimers and the probability P2 that the dimers grow to macroscopically large clusters before decay-ing to monomers. For that reason, when the supersaturation s is varied at constant temperature, available P2共s兲 data allow a model-independent evaluation of the nucleus size with the aid of Eqs.共10兲 or 共11兲.

The P2共s兲 dependence itself is relatively easily acces-sible by computer simulations of nucleation. Obtaining this dependence in such a way and using it in Eq. 共6兲 makes it possible to determine J in a rather wide s range. This new method of direct determination of J has the advantage to be independent of the cluster definition employed in the simu-lations.

The performed kinetic Monte Carlo simulations of 2D nucleation of monolayers on the共100兲 face of Kossel crystal demonstrate the potentialities of the new method: J is deter-mined over nearly 11 orders of magnitude and the corre-sponding 2D nucleus is constituted of about 1 to 30 mol-ecules. The classical nucleation theory overestimates the simulation nucleation rate by an s-dependent factor, and a recent corrected formula for J, Eq.共27兲, improves only quali-tatively the classical one, Eq.共25兲. Though describing rather

well the simulation findings for the nucleus size, the classical Gibbs-Thomson Eq.共22兲 is less successful than its corrected version, Eq.共30兲.

ACKNOWLEDGMENT

The authors gratefully acknowledge NWO共The Nether-lands Organization for Scientific Research兲 for a grant en-abling one of the authors共D.K.兲 to visit the Delft University of Technology and take part in the present work.

1F. F. Abraham, Homogeneous Nucleation Theory共Academic, New York,

1974兲.

2D. Kashchiev, Nucleation: Basic Theory with Applications

共Butterworth-Heinemann, Oxford, 2000兲.

3R. Becker and W. Döring, Ann. Phys. 24, 719共1935兲. 4G. M. White, J. Chem. Phys. 50, 4672共1969兲.

5J. H. ter Horst and D. Kashchiev, J. Chem. Phys. 119, 2241共2003兲. 6S. Karlin, A First Course in Stochastic Processes共Academic, New York,

1966兲, p. 205.

7T. L. Hill, J. Chem. Phys. 36, 3182共1962兲.

8T. L. Hill, Thermodynamics of Small Systems共Dover, New York, 1994兲,

Pt. II, p. 167.

9D. Kashchiev, J. Chem. Phys. 76, 5098共1982兲.

10Y. Viisanen, R. Strey, and H. Reiss, J. Chem. Phys. 99, 4680共1993兲. 11D. W. Oxtoby and D. Kashchiev, J. Chem. Phys. 100, 7665共1994兲. 12M. Volmer and A. Weber, Z. Phys. Chem., Stoechiom. Verwandtschaftsl.

119, 277共1926兲.

13J. H. ter Horst and P. J. Jansens, Surf. Sci. 574, 77共2005兲.

14D. Kashchiev, J. P. van der Eerden, and C. van Leeuwen, J. Cryst.

Growth 40, 47共1977兲.

15J. D. Weeks and G. H. Gilmer, Adv. Chem. Phys. 40, 157共1979兲. 16D. Kashchiev, J. Cryst. Growth 267, 685共2004兲.

17K. Binder, in Phase Transformations in Materials, edited by G. Kostorz

共Wiley, New York, 2001兲, p. 409.

18V. Bostanov, W. Obretenov, G. Staikov, and E. Budevski, J. Electroanal.

Chem. Interfacial Electrochem. 146, 303共1983兲.

19H. J. Leamy, G. H. Gilmer, and K. A. Jackson, in Surface Physics of

Materials, edited by J. B. Blakeley共Academic, New York, 1975兲, Vol. I,

p. 121.

20V. Talanquer, J. Chem. Phys. 106, 9957共1997兲. 21D. Kashchiev, J. Chem. Phys. 118, 1837共2003兲. 22I. Kusaka, J. Chem. Phys. 119, 1808共2003兲. 23D. Kashchiev, J. Chem. Phys. 120, 3749共2004兲. 24V. I. Kalikmanov, J. Chem. Phys. 121, 8916共2004兲.

Cytaty

Powiązane dokumenty

Obtained results yield the polar derivative analogues of some inequalities giving estimates for the growth of derivative of lacunary polynomials.. The Bernstein inequality that

As an illustration parts of the thalli of Parmelia conspersa (No. 16a) and Parmelia prolixa (No. 18) were taken and the mean value per surface unit of the thallus of both species

Ci i owi pisa- rze mogli co prawda za Gomułki, czy łaskawiej panującego Gierka, na zasadzie wyłączności oraz kontrolowanego odgórnie pozwolenia naruszać niekiedy obszar tabu

Przede wszystkim jednak zaznaczają, że w ostatnich latach badania naukowe nabrały istotnego znaczenia dla Kolumbii, a szczególnie dla ko- lumbijskich uniwersytetów, które

O dm iany pism a, które jeszcze nie odnosiły się do dźwięków języka, a jedynie do idei czy pojęć, określane są jako proto-pism a (piktogram y, ideogram y), n

Bogaty materiał faktograficzny przedstawiony przez autora prowadzi do wniosku, że za znany­ mi z dotychczasowych publikacji rozmowami politycznymi między emigracyjnymi organami

У контексті цього дослідження такий висновок дозволяє говорити про те, що, з огляду на відсутність у Законі України «Про авторське право

For functions of one and two real variables, the rate of pointwise and uniform convergence of some Bernstein type algebraic polynomials is investigated..