• Tidak ada hasil yang ditemukan

Directory UMM :wiley:Public:journals:jcc:suppmat:20:

N/A
N/A
Protected

Academic year: 2017

Membagikan "Directory UMM :wiley:Public:journals:jcc:suppmat:20:"

Copied!
15
0
0

Teks penuh

(1)

Supplementary Material for:

MMFF VII. Characterization of MMFF94, MMFF94s, and Other

Widely Available Force Fields for Conformational Energies and for

Intermolecular-Interaction Energies and Geometries.

THOMAS A. HALGREN

Molecular Systems, Merck Research Laboratories, P.O. Box 2000, Rahway, NJ 07065

Appendix A: Experimental Conformational Energies for Set 1

Table A.I lists some, but by no means all, of the experimental conformational energies that

are available for the comparisons in Set 1. It also lists experimental uncertainties (a few of

which we estimated from data in the original report), indicates the type of quantity measured,

and specifies the method used. When two or more experimental values are listed, the first is

used in the comparisons of conformational energies made in this paper. As will be noted, in a

few instances we have chosen a different experimental value than the one we employed in our

earlier work.1 We also wish to point out that our earlier paper cited previous theoretical papers

as the basis for the listed values, whereas we now give citations to the original literature. For

conformational Set 2, we adopted the experimental values cited by Gundertofte2 without further

examination.

The table shows that the measured conformational energies form a disparate set. In

particular, many different experimental techniques have been used. More significantly, the

derived quantities sometimes refers to the gas phase and sometimes to the liquid phase or to

(2)

many incorrect experimental results. This fact is established simply by noting the many

instances in which alternative experimental values for the same conformational difference differ

by many multiples of the stated uncertainty in either. In such cases, both experimental results

cannot be correct (except possibly where strong medium effects are involved).

Given that some of the listed experimental values are wrong, it is fair to ask how a choice

can be made. One approach is to consider the reliability of the experimental technique. Our

own, somewhat untutored, prejudice is for determinations based on spectroscopic intensities

measured as a function of temperature, either directly (indicated by "/T" in the table) or on

samples cryogenically deposited from different initial temperatures (indicated by "/cdT"). In

such cases, a plot of the logarithm of relative intensity vs. 1/T gives ∆Η directly. Even this

technique, however, can give poor results, as the differences of more than 1 kcal/mol between

alternative experimental values determined in this way for 1,difluoroethane and for

2-butanone attest. We also favor determinations in which vibrational band progressions are used

to construct a complete torsional potential along a one-dimensional torsional coordinate. In

these cases, differences between local minima then give ∆E for that coordinate (though we call

this quantity ∆Η0 in recognition of the fact that the subsumed coordinates include zero-point and

thermal vibrational energy contributions), in close analogy to energy differences taken between

local minima on a force field or quantum mechanical energy surface. Conversely,

determinations based on conformationally averaged nmr chemical shifts or coupling constants

for which the isolated-conformer end-point values could not be measured directly seem to us

particularly prone to error.

The date of the measurement is also a factor, as we presume that more recent values are more

likely to be accurate, in part because of improved experimental technique but also because any

substantial variation from previously published values is likely to have been closely examined

by the authors. In some cases, however, our choice simply came down to which measurement

(3)

paper. Given their general success in predicting the observed conformational energies, we

presume that these very high level quantum calculations are unlikely to be seriously in error in

any individual case. To be sure, this choice minimizes the rms "error" reported for GVB-LMP2

(cf. Table IV), and its use might seem to compromise our claim that this method is the most

accurate. In point of fact, however, even selection of the conformational energies from Table

A.I in worst agreement with the GVB-LMP2 results still finds that it best reproduces the

experimental data. Moreover, we can think of no better approach for dealing with cases in

which discordant experimental values appear to have equal claims to validity. We had used

much the same approach previously,1 but with "MP4SDQ(T)/TZP" calculations serving as the

best reference values then available to us.

The reasons underlying most choices reflected in the table should be apparent from the

foregoing. Thus, we shall comment here only on a few cases. One of these is 2-butanone,

where all the higher-level calculations in Table III, as well as several others we had previously

reported,1 clearly favor the lower, liquid-phase result. A second case is 1,3-butadiene, where

we had cited a gauche - trans conformational energy difference of 2.5 kcal/mol, this being the

average of values referenced by Wiberg et al.4 This value, as it happens, also corresponds to an

often cited2,3b,5 but suspect early determination in which Carreira constructed a one-dimensional

torsional potential but used an erroneous torsional coordinate that incorrectly identified the

second, higher-energy conformer as being cis rather than gauche.22 The higher, more recent

values20,21 shown in Table A.I seem to us preferable. They are also consistent with the pattern

in Table III in which the theoretical method finds a larger gauche - trans difference for

1,3-butadiene than for 2-methyl-1,3-1,3-butadiene, evidently because the more compact trans conformer

in the methyl compound is somewhat destabilized by nonbonded repulsions involving the added

2-methyl group; Table I shows that MMFF94 and most of the other force-field methods show

(4)

In the case of 1-butene, we had previously cited a conformational energy difference of 0.53

kcal/mol.1 We now realize, however, that this value is derived from the free-energy difference

of 0.94 kcal/mol measured by Van Hemelrijk et al.65 by simply applying an entropic correction

of R ln2 = 1.18 e.u. to allow for the two-fold statistical factor favoring skew. However, the

looser skew conformation is also expected to have a higher vibrational entropy. With this mind,

we have applied the full entropic correction of 2.3 e.u. favoring skew found by Durig and

Compton64 to arrived at the value of 0.24 ± 0.42 kcal/mol listed in Table A.I. This correction to

the measured ∆G brings the derived value for ∆Η into line with Durig and Compton's value of

0.22 ± 0.02 kcal/mol and with the GVB-LMP2 result for ∆E.

The conformational energy difference for methyl vinyl ether is of particular interest. We

ourselves had previously referenced the Sullivan, Dickson, and Durig value of 1.70 kcal/mol

favoring cis over skew.49 Friesner and co-workers,3b however, cited the earlier Durig and

Compton difference of 1.15 kcal/mol,51 and worked intently to improve their theoretical

methodology to reduce the large discrepancy with their higher LMP2/cc-pVTZ(-f) theoretical

estimate of 2.62 kcal/mol.6 They subsequently published a GVB-LMP2 result of 1.45 kcal/mol

favoring skew.3b However, calculations using the current version of Jaguar now find the

conformational energy difference to be 2.14 kcal/mol (Table III, manuscript) when a rigorously

consistent protocol for lone-pair delocalization is employed in the localized MP2 calculation in

connection with improved grid and dealiasing functions. Furthermore, we find that this value

falls to 1.61 kcal/mol when zero-point energy differences3a are taken into account. This last

estimate is fully consistent with the Sullivan, Dickson, and Durig result and with the enthalpy

difference of 1.59 kcal/mol obtained Beech et al.,50 the experimental result that caused Durig to

reinvestigate this system and led him to raise his lower early estimate.

For the "simple" but conformationally critical gauche - anti conformational difference in n

-butane, we list what now appears to be a definitive value of 0.67 ± 0.10 kcal/mol.55 Also shown

(5)

0.75 kcal/mol value we had previously cited1 on the basis of Allinger, Yuh, and Li's

recommendation.7 Much as in the case of methyl vinyl ether, both the new value and one of the

earlier values it has superseded are by Durig and co-workers. Unfortunately, Durig et al. do not

discuss the reasons for rejecting their higher earlier value. However, they do cite a number of

other experimental determinations, made using a variety of methods, that agree with their lower

value of 0.67 kcal/mol.55 Other recent high-level quantum calculations8 also support this value,

as does the GVB-LMP2/cc-pVTZ(-f) value of 0.72 kcal/mol cited in Table III of the

manuscript.

As an indication of the limitations of experimental data, we point out that the cis - trans

difference for methyl acetate is particularly tenuous: the listed ∆G of 8.5 ± 1 kcal/mol represents

the relative intensity of a single weak band in the infra-red carbonyl region that was observed in

a sample cryogenically deposited from a temperature of 879 K and was thought to probably

arise from the cis conformer.14 Indeed, the original authors cite this conformational energy

difference as being a ∆Η, though this interpretation apparently assumes no entropy difference

between cis and trans conformers. We think it remarkable that the theoretical values in Table

III of the manuscript agree with the experimental value as well as they do, but caution that a

even a disagreement of 1 – 2 kcal/mol should not be viewed as indicative of a problem with the

theoretical calculation.

Finally, we discuss two entries for which we had previously based our comparisons on

free-energy differences measured in solution. First, we now list a single value of ∆G = 1.15

kcal/mol for the axial - equatorial difference in cyclohexyl amine rather than the previously

cited range of 1.1 - 1.8 kcal/mol.1 The reason for this change is that the larger values in the

range are for determinations in polar media that are inappropriate for comparison to calculated

isolated-molecule properties. Even this new reference value, however, is higher than the

theoretical calculations cited in Table III can accommodate. Second, we previously cited a free

(6)

enthalpy difference of 0.58 kcal/mol found by Eliel and Gilbert in dilute nonpolar medium for

the model 4-t-butylcyclohexanol system;46 these authors criticized the consensus value of 0.52

kcal/mol that Hirsch recommended for nonpolar solvents47 as being too low because it

incorporates low values for early determinations that they believed to be unreliable.

Most other minor differences between the present data and that cited previously1 either reflect

the substitution of representative specific determinations for previously referenced average

(7)

TABLE A.I.

___________________________________________________

References for Experimental Conformational Energies (Set 1), in kcal/mol.

Conf. Comparisona Expt.b Deltac Typed Methode Ref.

N-methylformamide, c - t 1.4 ? ∆Ηsol ir/cdT,

nmr/T

9

N-methylacetamide, c - t 2.3

2.8

methyl formate, t - c 4.75

3.85

propionaldehyde, sk - cis 0.67

1.03

2-butanone, skew - cis 1.07

2.02

1,3-butadiene, g - s-trans 2.89

2.75

acrolein, cis - trans 1.70

1.89

1,2-difluoroethane, a - g 0.56

0.80

1,2-dichloroethane, g - a 1.08

1.09

1-fluoropropane, a - g 0.35

(8)

1-chloropropane, g - a 0.09

ethanol, gauche - anti 0.12

0.3, 0.5

butane, gauche - anti 0.67

0.97

(9)

2,3-dimethylbutane,

1-butene, cis - skew 0.22

0.07

2-butene, cis - trans 1.2

1.0

a In the conformational notation, idealized dihedral angles are: cis or c, 0°; trans, t, anti, or a,

180°; gauche or g, 60°; and skew or sk, 120°.

b When more than one value for the conformational energy is cited, the first listed value is used

in the comparisons in the body of this paper

c Error in measurement as cited by the authors or as explained in a note accompanying the

literature citation.

d gp = gas phase; sol = solution or liquid-phase; superscript "0" in ∆Η0 denotes a difference

between the bottoms of potential wells constructed for a one-dimensional torsional coordinate, lT

means determined at low temperature.

e "r", "ir", "mw", "uv", nmr, and "ed" denote raman, infrared, microwave, ultraviolet, nuclear

magnetic resonance, and electron-diffraction spectroscopy; "/T" means as "a function of

temperature"; "/cdT" indicates sample collection via cryogenic deposition from thermalized

molecular beams as a function of temperature; "/tp" signifies construction of a one-dimensional

torsional potential; "mCal" stands for microcalorimetry; "nmr/cc" and "nmr/cs" denote nmr

determinations based on conformationally averaged chemical shifts and coupling constants,

respectively; "ir/td" stands for temperature-drift ir spectroscopy; and "comb." indicates

(10)

1. T. A. Halgren and R. B. Nachbar, J. Comput. Chem., 17, 587-615 (1996).

2. K. Gundertofte, T. Liljefors, P.-O. Norrby, and I. Petterssen, J. Comput Chem., 17, 429-449

(1996).

3. (a) R. A. Friesner, R. B. Murphy, M. D. Beachy, M. N. Ringnalda, W. T. Pollard, B. D.

Duneitz, and Y. Cao, J. Chem. Phys., submitted (1998); (b) R. B. Murphy, W. T. Pollard, and R.

A. Friesner, J. Chem. Phys., 106, 5073-5084 (1997).

4. K. B. Wiberg, P. R. Rablen, and M. Marquez, J. Am. Chem. Soc., 114, 8654-8668 (1992).

5. (a) N. L. Allinger, F. Li, L. Yan, and J. C. Tai, J. Comput. Chem., 11, 868-895 (1990); (b) C.

J. Casewit, K. S. Colwell, and A. K. Rappé, J. Am. Chem. Soc., 114, 10035-100046 (1992).

6. R. B. Murphy, M. D. Beachy, R. A. Friesner, and M. N. Ringnalda, J. Chem. Phys., 103,

1481-1490 (1995).

7. N. L. Allinger, Y. H. Yuh, and J.-H. Lii, J. Am Chem. Soc., 111, 8551-8566 (1989). See also

N. L. Allinger and L. Yan, J. Am. Chem. Soc., 115, 11918-11925 (1993), and references therein.

8. G. D. Smith and R. L. Jaffe. J. Phys. Chem., 100, 18718-18724 (1996).

9. Average of values in: S. Ataka, H. Takeuchi, and M. Tasumi, J. Mol. Struct., 113, 147-160

(1984), and F. A. L. Anet and M. Squillacote, J. Am. Chem. Soc., 97, 3243-3244 (1974).

10. S. Ataka, H. Takeuchi, and M. Tasumi, J. Mol. Struct., 113, 147-160 (1984).

11

.

T. Drakenberg and S. Forsén, Chem. Commun., 1404-1405 (1971). This result is from a

difference in enthalpies of activation for isomerization in C2H4Cl2 as determined by nmr

line-shape analysis.

12. W. H. Hocking, Naturforsch., 31a, 1113-1121 (1976).

13. B. P. van Eijck and F. B. van Duijneveldt, J. Mol. Struct., 39, 157-163 (1977).

(11)

15. S. Ruschin and S. H. Bauer, J. Phys. Chem., 84, 3061-3064 (1980).

16. J. M. Riveros and E. B. Wilson Jr., J. Chem. Phys., 46, 4605-4612 (1967). This quantity is

reported to be the "energy difference" between the lowest vibrational levels, but its calculation

from measured microwave intensities at a single temperature suggests that it may be a

free-energy difference, presumably corrected for the two-fold statistical factor favoring the C-O-C-C

gauche conformer; no later report on this measurement appears to have been made.

17. J. R. Durig, D. A. C. Compton, and A. Q. McArver, J. Chem. Phys., 73, 719-724 (1980).

18. G. Sbrana and V. Schettino, J. Molec. Spectrosc., 33, 100-108 (1970). This value contributes

to the composite value of 0.95 kcal/mol cited in ref. 1. For other references, see: P. van Nuffel,

L. van den Enden, C. van Alsenoy, and H. J. Geise, J. Mol. Struct., 116, 99-118 (1984).

19. T. Shimanouchi, Y. Abe, and M. Makami, Spectrochim. Acta A, 24A, 1037-1053 (1968).

20. Y. N. Panchenko, A. V. Abramenkov and C. W. Bock, J. Mol. Struct., 140, 87-92 (1985).

21. J. R. Durig, W. E. Bucy, and A. R. H. Cole, Can. J. Phys., 53, 1832-1837 (1975).

22. L. Carreira, J. Chem. Phys.,62, 3851-3854 (1975).

23. Y. N. Panchenko, V. I. Pupyshev, A. V. Abramenkov, M. Traetteberg, and S. J. Cyvin, J.

Mol. Struct., 130, 355-359 (1985).

24. C. E. Blom and A. Bauder, Chem. Phys. Lett., 88, 55-58 (1982). Note that this value

supersedes the earlier result of 1.6 kcal/mol obtained by C. E. Blom, R. P. Müller and Hs. H.

Günthard, Chem. Phys. Lett., 73, 483-486 (1980).

25. E. J. Bair, W. Goetz, and D. A. Ramsay, Can. J. Phys., 49, 2710-2717 (1971). A note added

in proof to this paper reports a correction, agreed to by the original authors, to the value of 2.00 ±

0.11 kcal/mol published by: A. C. P. Alves, J. Christoffersen, and J. M. Hollas, Mol. Phys., 20,

(12)

26. P. Huber-Wälchi and Hs. H. Günthard, Chem. Phys. Lett., 30, 347-351. This result and the

cited uncertainty were obtained from the relative intensity vs. 1/T plots in Fig. 3. The authors

appear to incorrectly label their determinations as ∆G.

27. K. T. Hirano, S. Nonoyama, T. Miyajima, Y. Kurita, T. Kawamura, and H. Sato, J. Chem.

Soc., Chem. Commun., 1986, 606-607.

28. W. C. Harris, J. R. Holtzclaw, and V. F. Kalasinsky, J. Chem. Phys., 67, 3330-3338 (1977).

29. Average of values in Table 5 of: K. Kveseth, Acta Chem. Scand. A, 29, 307-311 (1975).

30. K. Tanabe, Spectrochim. Acta, 28A, 407-424 (1972). An entropy difference calculated from

symmetry numbers, moments of inertia, and vibrational frequencies was used to correct the

measured free-energy difference.

31. J. R. Durig, S. E. Godbey, and J. F. Sullivan, J. Chem. Phys., 80, 5983-5993 (1984).

32. E. Hirota, J. Chem. Phys., 37, 283-291 (1962).

33. W. A. Herrebout and B. J. van der Veken, J. Phys. Chem., 100, 9671-9677 (1996). I thank

Dr. Michael Beachy (Schrodinger) for supplying this reference.

34. K. Yamanouchi, M. Sugie, H. Takeo, C. Matsumura, and K. Kuchitsu, J. Phys. Chem., 88,

2315-2323 (1984). An entropy difference calculated from moments of inertia and vibrational

frequencies was used to correct the measured free-energy difference of -0.12 kcal/mol.

35. A. J. de Hoog, H. R. Buys, C. Altona, and E. Havinga, Tetrahedron, 25, 3365-3375 (1969).

Position of the equilibrium as a function of temperature was determined from conformationally

averaged nmr coupling constants.

36. R. M. Clay, G. M. Kelle, and F. G. Riddell, J. Am. Chem. Soc., 95, 4632 (1973).

37. J. R. Durig, G. A. Guirgis, and D. A. C. Compton, J. Phys. Chem., 83, 1313-1323 (1979).

(13)

39. Average of values for nonpolar solvents cited in H. Booth and M. L. Jozefowicz, J. Chem.

Soc., Perkins Trans. II, 895-901 (1976).

40. R. W. Baldock and A. R. Katritzky, J. Chem. Soc. B, 1470-1477 (1968); see also R. A. Y.

Jones, A. R. Katritzky, A. C. Richards, R. J. Wyatt, R. J. Bishop, and L. E. Sutton, J. Chem. Soc.

B, 127-131 (1970).

41. P. J. Crowley, M. J. T. Robinson, and M. G. Ward, Tetrahedron, 33, 915-925 (1977).

Kinetically-controlled protonation was used to trap conformers in quaternary salt form for nmr

analysis.

42. R. K. Kakar, and C. R. Quade, J. Chem. Phys., 72, 4300-4307 (1980).

43. J. R. Durig, W. E. Bucy, C. J. Wurrey, and L. A. Carriera, J. Phys. Chem., 79, 988-993

(1975). The authors note that the value chosen for ∆H depends on the choice for a key line

assignment.

44. H. L. Fang and R. L. Swofford, Chem. Phys. Lett., 105, 5-11 (1984). Signal detection was by

photoacoustic spectroscopy.

45. E. Hirota, as cited in W. A. Lathan, L. Radom, W. J. Hehre, and J. A. Pople, J. Am. Chem.

Soc., 95, 699-703 (1973). The reference to relative microwave intensities with no reference to

temperature dependence suggests that this measurement is ∆G, presumably corrected for the

two-fold statistical factor favoring the gauche conformer. For the preliminary study, see: S. Kondo

and E. Hirota, J. Mol. Spectrosc.,34, 97-107 1970). No follow-up study seems to have appeared.

46. E. L. Eliel and E. C. Gilbert, J. Am. Chem. Soc., 91, 5487-5495 (1969). The cited value was

measured for 4-t-butylcyclohexanol epimers equilibrated via Raney nickel catalysis.

47. J. A. Hirsch, in N. L. Allinger and E. L. Eliel, Eds., Topics in Stereochemistry, 1, 199-222

(1967). This often-cited value (cf. refs. 1, 2, 3b, and 5) is the free-energy difference Hirsch

recommends for nonpolar solvents, but has been criticized by Eliel and Gilbert46 as being too

(14)

48. T. Kitagawa and T. Miyazawa, Bull. Chem. Soc., Jpn., 41, 1976 (1968).

49. J. F. Sullivan, T. J. Dickson, and J. R. Durig, Spectrochim. Acta, 42A, 113-122 (1986).

50. T. Beech, R. Gunde, P. Felder, and Hs. Günthard, Spectrochim. Acta., 41A, 319-339 (1985).

51. J. R. Durig and D. A. C. Compton, J. Chem. Phys., 69, 2028-2935 (1978). Note that this

value is incorrect and has been superseded by the Sullivan, Dickson and Durig result.49

52. H. Weiser, W. G. Laidlaw, P. J. Kruger, and H. Fuhrer, Spectrochim. Acta, 24A, 1055-1089

(1968). The listed uncertainty of ± 0.05 reflects a purely subjective assessment on our part of the

data in Fig. 13 of this reference, and if anything is likely to be an overestimate.

53. F. R. Jensen, C. H. Bushweller, and B. H. Beck, J. Am. Chem. Soc.,91, 344-351 (1969).

54. E. L. Eliel and E. C. Gilbert, J. Am. Chem. Soc., 91, 5487-5495 (1969). This value was

measured for 4-t-butylmethoxycyclohexane using conformationally averaged nmr chemical

shifts.

55. W. A. Herrebout, B. J. van der Veken, A. Wang, and J. R. Durig, J. Phys. Chem., 99,

578-585 (1995). I thank Prof. Alex MacKerell (Univ. Maryland) for supplying this reference.

56. A. L. Verma, W. F. Murphy, and H. J. Bernstein, J. Chem. Phys., 60, 1540-1544 (1974).

57. J. R. Durig, A. Wang, W. Beshir, and T. S. Little, J. Raman. Spec., 22, 683-704 (1991).

58. R. K. Heenan and L. S. Bartell, J. Chem. Phys., 78, 1270-1274 (1983); W. F. Bradford, S.

Fitzwater, and L. S. Bartell, J. Molec. Struct., 38, 185-194 (1977). These papers are cited by

Allinger et al.7 as the basis for the value and the uncertainty listed in this table.

59. M. Squillacote, R. S. Sheridan, O. L. Chapman, and F. A. L. Anet, J. Am. Chem. Soc., 97,

3244-3246 (1975); this result uses data from: F. A. L. Anet and A. J. Bourn, J. Am. Chem. Soc.,

89, 760-768 (1967).

(15)

61. T. L. Boates, thesis, Iowa State University, 1966, as cited by: E. J. Jacob, H. B. Thompson,

and L. S. Bartell, J. Chem. Phys., 47, 3736-3753. This value is reported as a ∆G, but has been

corrected for the two-fold statistical factor in favor of the gauche conformer.

62. F. A. L. Anet and V. J. Basus, J. Am. Chem. Soc., 95, 4424-4426 (1973).

63. F. A. L. Anet and J. Krane, Isr. J. Chem., 20, 72-83 (1980). This result was measured by

nmr line-shape analysis at -145° and converted from ∆G = 0.5 kcal/mol by using a

force-field-calculated ∆S of 3.5 e.u. Use of the MP2/6-31G* estimate of 4.33 e.u. for ∆S (cf. ref. 1) would

raise ∆H by an additional 0.1 kcal/mol.

64. J. R. Durig and D. A. C. Compton, J. Phys. Chem., 84, 773-781 (1980). We estimated the

listed uncertainty of ± 0.02 kcal/mol from hand-drawn slopes for the five highest vs. four lowest

temperature points in Fig. 3 of this reference.

65. D. Van Hemelrijk, L. Van den Enden, H. J. Greise, H. L. Sellers, and L. Schäfer, J. Am.

Chem. Soc., 102, 2189-2195 (1980). The listed ∆H differs from the reported ∆H of 0.53

kcal/mol because it uses Durig and Compton's measured ∆S of 2.3 e.u. favoring skew (cf. ref. 64)

to make the full entropic correction (not just the symmetry-number correction of R ln2 = 1.18

e.u.) to the measured free-energy difference of 0.94 kcal/mol.

66. D. M. Golden, K. W. Egger, and S. W. Benson, J. Am. Chem. Soc., 86, 5416-5420 (1964).

67. Heat of combustion value cited in, but not correctly referenced by: N. L. Allinger, F. Li, and

L. Yan, J. Comput. Chem., 11, 848-867 (1990). The value listed here is from heats of formation

given in: Lange's Handbook of Chemistry, Fourteenth Edition, J. A. Dean, Ed., McGraw-Hill:

New York, NY, 1992. Measured heats of hydrogenation give similar values of 1.1 - 1.2

kcal/mol; see M. R. Ibrahim, Z. A. Fataftah, P. von R. Schleyer, and P. D. Stout, J. Comput.

Gambar

TABLE A.I. ___________________________________________________

Referensi

Dokumen terkait

Laporan kerja praktek yang berjudul “ IMPLEMENTASI SISTEM INFORMASI MANAJEMEN DAERAH BARANG MILIK DAERAH (SIMDA- BMD) PADA DINAS PENGELOLAAN KEUANGAN DAN ASET DAERAH

Piutang wesel yang timbul sebagai akibat penjualan barang atau jasa secara kredit akan dilaporkan dalam neraca sebagai aktiva lancar, sedangkan wesel yang timbul

These are times of exciting change and history shows that with changing social contexts, the nature of literacy and literacy learning is redeined. With the advent of new

[r]

Pekerjaan Pengoperasian Pesawat Lifts Gedung Utama (Berikut Penggantian Suku.. Cadang) Kantor Kementerian Perdagangan

Modal intelektual yang dimiliki semua dosen Universitas Komputer Indonesia umumnya sudah baik, dapat dilihat dari kemampuan dosen dalam memberikan materi yang mudah dimengerti

terutama mengenai parenkim paru dalam bentuk bronkopneumonia dan pneumonia lobar serta pneumonia bakterial tipe campuran atipikal yaitu perjalanan penyakit ringan

Pneumonia bakterial berupa : pneumonia bakterial tipe tipikal yang terutama mengenai parenkim paru dalam bentuk bronkopneumonia dan pneumonia lobar serta pneumonia bakterial