( 2 Jun 94)
 
                      ***********************************
                      *                                 *
                      * Section 4 - Further Information *
                      *                                 *
                      ***********************************
 
          This section of the manual contains both references, and
          hints on how to do things.  The following is a list of
          the topics covered:
 
             o  Computational References.
 
             o  Basis Set References, and descriptions.
 
             o  How to do RHF, ROHF, UHF, and GVB calculations.
                     General considerations
                     Direct SCF
                     SCF convergence accelerators
                     High spin open shell SCF (ROHF)
                     Other open shell SCF cases (GVB)
                     True GVB perfect pairing runs
                     The special case of TCSCF
                     A caution about symmetry with GVB
 
             o  How to do CI and MCSCF calculations.
                     Specification of the wavefunction
                     Use of symmetry
                     Selection of orbitals
                     The iterative process
                     Miscellaneous hints
                     MCSCF implementation
                     References
 
             o  Geometry Searches and Internal Coordinates.
                     Quasi-Newton searches
                     The hessian
                     Coordinates choices
                     The role of symmetry
                     Practical matters
                     Saddle points
                     Mode following
 
             o  MOPAC calculations within GAMESS
 
             o  Molecular Properties, and conversion factors.
 
             o  Localization tips.
 
             o  Intrinsic Reaction Coordinate (IRC) methods
 
             o  Transition moments and spin-orbit coupling.
 

 
                          Computational References
                          ------------- ----------
 
          GAMESS - 
             M.W.Schmidt, K.K.Baldridge, J.A.Boatz, S.T.Elbert,
             M.S.Gordon, J.J.Jensen, S.Koseki, N.Matsunaga, 
             K.A.Nguyen, S.Su, T.L.Windus, M.Dupuis, J.A.Montgomery
             J.Comput.Chem. 14, 1347-1363 (1993)

          HONDO -
          "The General Atomic and Molecular Electronic Structure
             System: HONDO 7.0"  M.Dupuis, J.D.Watts, H.O.Villar,
             G.J.B.Hurst  Comput.Phys.Comm. 52, 415-425(1989)
          "HONDO: A General Atomic and Molecular Electronic
             Structure System"  M.Dupuis, P.Mougenot, J.D.Watts,
             G.J.B.Hurst, H.O.Villar in "MOTECC: Modern Techniques
             in Computational Chemistry"  E.Clementi, Ed.
             ESCOM, Leiden, the Netherlands, 1989, pp 307-361.
          "HONDO: A General Atomic and Molecular Electronic
             Structure System"  M.Dupuis, A.Farazdel, S.P.Karna,
             S.A.Maluendes in "MOTECC: Modern Techniques in
             Computational Chemistry"  E.Clementi, Ed.
             ESCOM, Leiden, the Netherlands, 1990, pp 277-342.
          M.Dupuis, S.Chin, A.Marquez in Relativistic and Electron
          Correlation Effects in Molecules, G.Malli, Ed. Plenum 
          Press, NY 1994.

          The final three papers describes many of the algorithms
          in detail, and much of this paper applies also to GAMESS.
 
          sp integrals -
          J.A.Pople, W.J.Hehre  J.Comput.Phys. 27, 161-168(1978)
 
          sp gradient integrals -
          A.Komornicki, K.Ishida, K.Morokuma, R.Ditchfield,
          M.Conrad  Chem.Phys.Lett. 45, 595-602(1977)
 
          spdfg integrals -
          "Numerical Integration Using Rys Polynomials"
              H.F.King and M.Dupuis   J.Comput.Phys. 21,144(1976)
          "Evaluation of Molecular Integrals over Gaussian
                                               Basis Functions"
             M.Dupuis,J.Rys,H.F.King  J.Chem.Phys. 65,111-116(1976)
          "Molecular Symmetry and Closed Shell HF Calculations"
           M.Dupuis and H.F.King   Int.J.Quantum Chem. 11,613(1977)
          "Computation of Electron Repulsion Integrals using
                     the Rys Quadrature Method"
              J.Rys,M.Dupuis,H.F.King J.Comput.Chem. 4,154-157(1983)
 
          spd gradient integrals -
          "Molecular Symmetry. II. Gradient of Electronic Energy
           with respect to Nuclear Coordinates"
              M.Dupuis and H.F.King  J.Chem.Phys. 68,3998(1978)
 

 
          spd hessian integrals -
          "Molecular Symmetry. III. Second derivatives of Electronic
           Energy with respect to Nuclear Coordinates"
              T.Takada, M.Dupuis, H.F.King
              J.Chem.Phys. 75, 332-336 (1981)
 
          RHF -
          C.C.J. Roothaan      Rev.Mod.Phys. 23, 69(1951)
 
          UHF -
          J.A. Pople, R.K. Nesbet  J.Chem.Phys 22, 571 (1954)
 
          GVB and OCBSE-ROHF -
          F.W.Bobrowicz and W.A.Goddard, in Modern Theoretical
          Chemistry, Vol 3, H.F.Schaefer III, Ed., Chapter 4.
 
          ROHF -
          R.McWeeny, G.Diercksen J.Chem.Phys. 49,4852-4856(1968)
          M.F.Guest, V.R.Saunders, Mol.Phys. 28, 819-828(1974)
          J.S.Binkley, J.A.Pople, P.A.Dobosh
             Mol.Phys.  28, 1423-1429 (1974)
          E.R.Davidson  Chem.Phys.Lett.  21,565(1973)
          K.Faegri, R.Manne  Mol.Phys.  31,1037-1049(1976)
          H.Hsu, E.R.Davidson, and R.M.Pitzer
             J.Chem.Phys. 65,609(1976)
 
          Direct SCF -
          J.Almlof, K.Faegri, K.Korsell
             J.Comput.Chem. 3, 385-399 (1982)
          M.Haser, R.Ahlrichs
             J.Comput.Chem. 10, 104-111 (1989)
 
          GUGA CI -
          B.Brooks and H.F.Schaefer  J.Chem. Phys. 70,5092(1979)
          B.Brooks, W.Laidig, P.Saxe, N.Handy, and H.F.Schaefer,
             Physica Scripta 21,312(1980).
 
          2nd order Moller-Plesset -
          M.Dupuis, S.Chin, A.Marquez, see above
 
          MOPAC 6 -
          J.J.P.Stewart  J.Computer-Aided Molecular Design
          4, 1-105 (1990)
          References for parameters for individual atoms may be
          found on the printout from your runs.
 
          DIIS (Direct Inversion in the Iterative Subspace) -
          P.Pulay  J.Comput.Chem. 3, 556-560(1982)


 
          RHF/ROHF/TCSCF coupled perturbed Hartree Fock -
          "Single Configuration SCF Second Derivatives on a Cray"
              H.F.King, A.Komornicki in "Geometrical Derivatives of
              Energy Surfaces and Molecular Properties" P.Jorgensen
              J.Simons, Ed. D.Reidel, Dordrecht, 1986, pp 207-214.
          Y.Osamura, Y.Yamaguchi, D.J.Fox, M.A.Vincent, H.F.Schaefer
              J.Mol.Struct. 103, 183-186 (1983)
          M.Duran, Y.Yamaguchi, H.F.Schaefer
              J.Phys.Chem. 92, 3070-3075 (1988)
 
          Davidson eigenvector method -
          E.R.Davidson  J.Comput.Phys. 17,87(1975) and "Matrix
          Eigenvector Methods" p. 95 in "Methods in Computational
          Molecular Physics" ed. by G.H.F.Diercksen and S.Wilson
 
          EVVRSP in memory diagonalization -
          S.T.Elbert  Theoret.Chim.Acta  71,169-186(1987)
 
          Energy localization-
          C.Edmiston, K.Ruedenberg  Rev.Mod.Phys.  35, 457-465(1963).
 
          Boys localization-
          S.F.Boys, "Quantum Science of Atoms, Molecules, and Solids"
          P.O.Lowdin, Ed, Academic Press, NY, 1966, pp 253-262.
 
          Population localization -
          J.Pipek, P.Z.Mezey  J.Chem.Phys.  90, 4916(1989).
 
          Geometry optimization and saddle point location -
          J.Baker  J.Comput.Chem. 7, 385-395(1986).
          T.Helgaker  Chem.Phys.Lett. 182, 503-510(1991).
          P.Culot, G.Dive, V.H.Nguyen, J.M.Ghuysen
              Theoret.Chim.Acta  82, 189-205(1992).
 
          Vibrational Analysis in Cartesian coordinates -
          W.D.Gwinn  J.Chem.Phys.  55,477-481(1971)
 
          Normal coordinate decomposition analysis -
          J.A.Boatz and M.S.Gordon,
             J.Phys.Chem. 93, 1819-1826(1989).
 
          Modified Virtual Orbitals (MVOs) -
          C.W.Bauschlicher, Jr.  J.Chem.Phys.  72,880-885(1980)
 

 
          Effective core potentials (ECP) -
          L.R.Kahn, P.Baybutt, D.G.Truhlar
             J.Chem.Phys. 65, 3826-3853 (1976)
          See also the papers listed for SBK and HW basis sets.

          TCGMSG
          R.J.Harrison, Int.J.Quantum Chem. 40, 847-863(1991)

          Bond orders
          I.Mayer, Chem.Phys.Lett., 97,270-274(1983), 117,396(1985).
          I.Mayer, Theoret.Chim.Acta, 67,315-322(1985).
 

 
                           Basis Set References
                           ----- --- ----------
 
               An excellent review of the relationship between
          the atomic basis used, and the accuracy with which
          various molecular properties will be computed is:
          E.R.Davidson, D.Feller  Chem.Rev. 86, 681-696(1986).
 
          STO-NG      H-Ne        Ref. 1 and 2
                      Na-Ar,      Ref. 2 and 3 **
                      K,Ca,Ga-Kr  Ref. 4
                      Rb,Sr,In-Xe Ref. 5
                      Sc-Zn,Y-Cd  Ref. 6
 
          1) W.J.Hehre, R.F.Stewart, J.A.Pople
             J.Chem.Phys. 51, 2657-2664(1969).
          2) W.J.Hehre, R.Ditchfield, R.F.Stewart, J.A.Pople
             J.Chem.Phys. 52, 2769-2773(1970).
          3) M.S.Gordon, M.D.Bjorke, F.J.Marsh, M.S.Korth
             J.Am.Chem.Soc. 100, 2670-2678(1978).
             ** the valence scale factors for Na-Cl are taken
                from this paper, rather than the "official"
                Pople values in Ref. 2.
          4) W.J.Pietro, B.A.Levi, W.J.Hehre, R.F.Stewart,
             Inorg.Chem. 19, 2225-2229(1980).
          5) W.J.Pietro, E.S.Blurock, R.F.Hout,Jr., W.J.Hehre, D.J.
             DeFrees, R.F.Stewart  Inorg.Chem. 20, 3650-3654(1980).
          6) W.J.Pietro, W.J.Hehre J.Comput.Chem. 4, 241-251(1983).
 
 
 
          MINI/MIDI    H-Xe       Ref. 9
 
          9) "Gaussian Basis Sets for Molecular Calculations"
             S.Huzinaga, J.Andzelm, M.Klobukowski, E.Radzio-Andzelm,
             Y.Sakai, H.Tatewaki   Elsevier, Amsterdam, 1984.
 
              The MINI bases are three gaussian expansions of each
          atomic orbital.  The exponents and contraction
          coefficients are optimized for each element, and s and p
          exponents are not constrained to be equal.  As a result
          these bases give much lower energies than does STO-3G.
          The valence MINI orbitals of main group elements are
          scaled by factors optimized by John Deisz at North Dakota
          State University.  Transition metal MINI bases are not
          scaled.  The MIDI bases are derived from the MINI sets by
          floating the outermost primitive in each valence orbitals,
          and renormalizing the remaining 2 gaussians.  MIDI bases
          are not scaled by GAMESS.  The transition metal bases are
          taken from the lowest SCF terms in the s**1,d**n
          configurations.
 

 
          3-21G       H-Ne           Ref. 10     (also 6-21G)
                      Na-Ar          Ref. 11     (also 6-21G)
          K,Ca,Ga-Kr,Rb,Sr,In-Xe     Ref. 12
                      Sc-Zn          Ref. 13
                      Y-Cd           Ref. 14
 
          10) J.S.Binkley, J.A.Pople, W.J.Hehre
              J.Am.Chem.Soc. 102, 939-947(1980).
          11) M.S.Gordon, J.S.Binkley, J.A.Pople, W.J.Pietro,
              W.J.Hehre  J.Am.Chem.Soc. 104, 2797-2803(1982).
          12) K.D.Dobbs, W.J.Hehre  J.Comput.Chem. 7,359-378(1986)
          13) K.D.Dobbs, W.J.Hehre  J.Comput.Chem. 8,861-879(1987)
          14) K.D.Dobbs, W.J.Hehre  J.Comput.Chem. 8,880-893(1987)
 
 
 
          N-31G   references for  4-31G         5-31G        6-31G
                      H            15            15           15
                      He           23            23           23
                      Li           19,24                      19
                      Be           20,24                      20
                      B            17                         19
                      C-F          15            16           16
                      Ne           23                         23
                      Na-Ga                                   22
                      Si                                      21 **
                      P-Cl         18                         22
                      Ar                                      22
 
          15) R.Ditchfield, W.J.Hehre, J.A.Pople
              J.Chem.Phys. 54, 724-728(1971).
          16) W.J.Hehre, R.Ditchfield, J.A.Pople
              J.Chem.Phys. 56, 2257-2261(1972).
          17) W.J.Hehre, J.A.Pople J.Chem.Phys. 56, 4233-4234(1972).
          18) W.J.Hehre, W.A.Lathan J.Chem.Phys. 56,5255-5257(1972).
          19) J.D.Dill, J.A.Pople J.Chem.Phys. 62, 2921-2923(1975).
          20) J.S.Binkley, J.A.Pople J.Chem.Phys. 66, 879-880(1977).
          21) M.S.Gordon  Chem.Phys.Lett. 76, 163-168(1980)
              ** - Note that the built in 6-31G basis for Si is
                   not that given by Pople in reference 22.
                   The Gordon basis gives a better wavefunction,
                   for a ROHF calculation in full atomic (Kh)
                   symmetry,
                   6-31G      Energy       virial
                   Gordon   -288.828573   1.999978
                   Pople    -288.828405   2.000280
                   See the input examples for how to run in Kh.
          22) M.M.Francl, W.J.Pietro, W.J.Hehre, J.S.Binkley,
              M.S.Gordon, D.J.DeFrees, J.A.Pople
              J.Chem.Phys. 77, 3654-3665(1982).
          23) Unpublished, copied out of GAUSSIAN82.
          24) For Li and Be, 4-31G is actually a 5-21G expansion.
 

 
          Extended basis sets
 
              6-311G
          28) R.Krishnan, J.S.Binkley, R.Seeger, J.A.Pople
              J.Chem.Phys. 72, 650-654(1980).
 
              valence double zeta "DZV" sets:
 
              "DH" basis - DZV for H, Li-Ne, Al-Ar
          30) T.H.Dunning, Jr., P.J.Hay  Chapter 1 in "Methods of
              Electronic Structure Theory", H.F.Shaefer III, Ed.
              Plenum Press, N.Y. 1977, pp 1-27.
              Note that GAMESS uses inner/outer scale factors of
              1.2 and 1.15 for DH's hydrogen (since at least 1983).
              To get Thom's usual basis, scaled 1.2 throughout:
                  HYDROGEN   1.0   x, y, z
                     DH  0  1.2   1.2
              "BC" basis - DZV for Ga-Kr
          31) R.C.Binning, Jr., L.A.Curtiss
              J.Comput.Chem. 11, 1206-1216(1990)
 
 
              valence triple zeta "TZV" sets:
 
              TZV for Li-Ne
          40) T.H. Dunning, J.Chem.Phys. 55 (1971) 716-723.
              TZV for Na-Ar - also known as the "MC" basis
          41) A.D.McLean, G.S.Chandler
              J.Chem.Phys. 72,5639-5648(1980).
              TZV for K,Ca
          42) A.J.H. Wachters, J.Chem.Phys. 52 (1970) 1033-1036.
              (see Table VI, Contraction 3).
              TZV for Sc-Zn (taken from HONDO 7)
          This is Wachters' (14s9p5d) basis (ref 42) contracted
          to (10s8p3d) with the following modifications
                 1. the most diffuse s removed;
                 2. additional s spanning 3s-4s region;
                 3. two additional p functions to describe the 4p;
                 4. (6d) contracted to (411) from ref 43,
                    except for Zn where Wachter's (5d)/[41]
                    and Hay's diffuse d are used.
          43) A.K. Rappe, T.A. Smedley, and W.A. Goddard III,
              J.Phys.Chem. 85 (1981) 2607-2611
 

 
 
          Valence only basis sets (for use with corresponding ECPs)
 
          SBK   -31G splits, bigger for trans. metals (available Li-Rn)
          50) W.J.Stevens, H.Basch, M.Krauss
                  J.Chem.Phys. 81, 6026-6033 (1984)
          51) W.J.Stevens, H.Basch, M.Krauss, P.Jasien
                  Can.J.Chem, 70, 612-630 (1992)
          52) T.R.Cundari, W.J.Stevens  
                  J.Chem.Phys. 98, 5555-5565(1993)
 
          HW    -21 splits (sp exponents not shared)
              transition metals (not built in at present, although
              they will work if you type them in).
          53) P.J.Hay, W.R.Wadt  J.Chem.Phys.  82, 270-283 (1985)
              main group (available Na-Xe)
          54) W.R.Wadt, P.J.Hay  J.Chem.Phys.  82, 284-298 (1985)
              see also
          55) P.J.Hay, W.R.Wadt  J.Chem.Phys.  82, 299-310 (1985)
 
 
          Polarization exponents
 
              STO-NG*
          60) J.B.Collins, P. von R. Schleyer, J.S.Binkley,
              J.A.Pople  J.Chem.Phys. 64, 5142-5151(1976).
 
              3-21G*.   See also reference 12.
          61) W.J.Pietro, M.M.Francl, W.J.Hehre, D.J.DeFrees,  J.A.
              Pople, J.S.Binkley J.Am.Chem.Soc. 104,5039-5048(1982)
 
              6-31G* and 6-31G**.   See also reference 22 above.
          62) P.C.Hariharan, J.A.Pople
              Theoret.Chim.Acta 28, 213-222(1973)
 

 
 
          STO-NG*  means d orbitals are used on third row atoms only.
                   The original paper (ref 60) suggested z=0.09 for
                   Na and Mg, and z=0.39 for Al-Cl.
                   At NDSU we prefer to use the same exponents as
                   in 3-21G* and 6-31G*, so we know we're looking
                   at changes in the sp basis, not the d exponent.
 
          3-21G*   means d orbitals on main group elements in the
                   third and higher periods.  Not defined for the
                   transition metals, where there are p's already
                   in the basis.  Except for alkalis and alkali
                   earths, the 4th and 5th row zetas are from
                   Huzinaga, et al (ref 9).  The exponents are
                   normally the same as for 6-31G*.
 
          6-31G*   means d orbitals on second and third row atoms.
                   We use Mark Gordon's z=0.395 for silicon, as well
                   as his fully optimized sp basis (ref 21).
                   This is often written 6-31G(d) today.
 
          6-31G**  means the same as 6-31G*, except that p functions
                   are added on hydrogens.
                   This is often written 6-31G(d,p) today.
 
          6-311G** means p orbitals on H, and d orbitals elsewhere.
                   The exponents were derived from correlated atomic
                   states, and so are considerably tighter than the
                   polarizing functions used in 6-31G**, etc.
                   This is often written 6-311G(d,p) today.
 
              The definitions for 6-31G* for C-F are disturbing in
          that they treat these atoms the same.  Dunning and Hay
          (ref 30) have recommended a better set of exponents for
          second row atoms and a slightly different value for H.
 
              2p, 3p, 2d, 3p polarization sets are usually thought
          of as arising from applying splitting factors to the
          1p and 1d values.  For example, SPLIT2=2.0, 0.5 means
          to double and halve the single value.  The default
          values for SPLIT2 and SPLIT3 are taken from reference
          72, and were derived with correlation in mind.  The
          SPLIT2 values often produce a higher (!) HF energy than
          the singly polarized run, because the exponents are
          split too widely.  SPLIT2=0.4,1.4 will always lower the
          SCF energy (the values are the unpublished personal
          preference of MWS), and for SPLIT3 we might suggest
          3.0,1.0,1/3.
 
              With all this as background, we are ready to present
          the table of polarization exponents built into GAMESS.
 

 
              Built in polarization exponents, chosen by POLAR=
          in the $BASIS group.  The values are for d functions
          unless otherwise indicated.
 
              Please note that the names associated with each
          column are only generally descriptive.  For example, the
          column marked "Pople" contains a value for Si with which
          John Pople would not agree, and that the K-Xe values in
          this column were actually originally from the Huzinaga
          group.  The exponents for Ga-Kr are from the Binning and
          Curtiss paper, not Thom Dunning.  And so on.
 
                 POPLE    POPN311   DUNNING   HUZINAGA    HONDO7
                 ------   -------   -------   --------    ------
            H    1.1(p)    0.75(p)   1.0(p)     1.0(p)    1.0(p)
            He   1.1(p)    0.75(p)   1.0(p)     1.0(p)    1.0(p)
 
            Li   0.2       0.200                0.076(p)
            Be   0.4       0.255                0.164(p)  0.32
            B    0.6       0.401     0.70       0.388     0.50
            C    0.8       0.626     0.75       0.600     0.72
            N    0.8       0.913     0.80       0.864     0.98
            O    0.8       1.292     0.85       1.154     1.28
            F    0.8       1.750     0.90       1.496     1.62
            Ne   0.8       2.304     1.00       1.888     2.00
 
            Na   0.175                          0.061(p)  0.157
            Mg   0.175                          0.101(p)  0.234
            Al   0.325                          0.198     0.311
            Si   0.395                          0.262     0.388
            P    0.55                           0.340     0.465
            S    0.65                           0.421     0.542
            Cl   0.75                           0.514     0.619
            Ar   0.85                           0.617     0.696
 
            K    0.1                            0.039(p)
            Ca   0.1                            0.059(p)
            Ga   0.207               0.141
            Ge   0.246               0.202
            As   0.293               0.273
            Se   0.338               0.315
            Br   0.389               0.338
            Kr   0.443               0.318
 
            Rb   0.11                           0.034(p)
            Sr   0.11                           0.048(p)
 
          A blank means the value equals the "Pople" column.
 
          Common d polarization for all sets:
              In     Sn     Sb     Te      I     Xe
            0.160  0.183  0.211  0.237  0.266  0.297
              Tl     Pb     Bi     Po     At     Rn
            0.146  0.164  0.185  0.204  0.225  0.247
 

 
          Anion diffuse functions
 
              3-21+G, 3-21++G, etc.
          70) T.Clark, J.Chandrasekhar, G.W.Spitznagel, P. von R.
              Schleyer J.Comput.Chem. 4, 294-301(1983)
          71) G.W.Spitznagel, Diplomarbeit, Erlangen, 1982.
          72) M.J.Frisch, J.A.Pople, J.S.Binkley J.Chem.Phys.
              80, 3265-3269 (1984).
 
              Anions usually require diffuse basis functions to
          properly represent their spatial diffuseness.  The use of
          diffuse sp shells on atoms in the second and third rows is
          denoted by a + sign, also adding diffuse s functions on
          hydrogen is symbolized by ++.  These designations can be
          applied to any of the Pople bases, e.g.  3-21+G, 3-21+G*,
          6-31++G**.  The following exponents are for L shells,
          except for H.  For H-F, they are taken from ref 70.  For
          Na-Cl, they are taken directly from reference 71.  These
          values may be found in footnote 13 of reference 72.
          For Ga-Br, In-I, and Tl-At these were optimized for the
          atomic ground state anion, using ROHF with a flexible ECP
          basis set, by Ted Packwood at NDSU.
 
              H
           0.0360
             Li      Be       B       C       N       O       F
           0.0074  0.0207  0.0315  0.0438  0.0639  0.0845  0.1076
             Na      Mg      Al      Si       P       S      Cl
           0.0076  0.0146  0.0318  0.0331  0.0348  0.0405  0.0483
                             Ga      Ge      As      Se      Br
                           0.0205  0.0222  0.0287  0.0318  0.0376
                             In      Sn      Sb      Te       I
                           0.0223  0.0231  0.0259  0.0306  0.0368
                             Tl      Pb      Bi      Po      At
                           0.0170  0.0171  0.0215  0.0230  0.0294
 
          Additional information about diffuse functions and also
          Rydberg type exponents can be found in reference 30.
 

 
              The following atomic energies are from UHF
          calculations (RHF on 1-S states), with p orbitals not
          symmetry equivalenced, and using the default molecular
          scale factors.  They should be useful in picking a basis
          of the desired energy accuracy, and estimating the correct
          molecular total energies.
 
 
          Atom state   STO-2G        STO-3G       3-21G       6-31G
          H   2-S     -.454397     -.466582     -.496199    -.498233
          He  1-S    -2.702157    -2.807784    -2.835680   -2.855160
          Li  2-S    -7.070809    -7.315526    -7.381513   -7.431236
          Be  1-S   -13.890237   -14.351880   -14.486820  -14.566764
          B   2-P   -23.395284   -24.148989   -24.389762  -24.519492
          C   3-P   -36.060274   -37.198393   -37.481070  -37.677837
          N   4-S   -53.093007   -53.719010   -54.105390  -54.385008
          O   3-P   -71.572305   -73.804150   -74.393657  -74.780310
          F   2-P   -95.015084   -97.986505   -98.845009  -99.360860
          Ne  1-S  -122.360485  -126.132546  -127.803825 -128.473877
          Na  2-S  -155.170019  -159.797148  -160.854065 -161.841425
          Mg  1-S  -191.507082  -197.185978  -198.468103 -199.595219
          Al  2-P  -233.199965  -239.026471  -240.551046 -241.854186
          Si  3-P  -277.506857  -285.563052  -287.344431 -288.828598
          P   4-S  -327.564244  -336.944863  -339.000079 -340.689008
          S   3-P  -382.375012  -393.178951  -395.551336 -397.471414
          Cl  2-P  -442.206260  -454.546015  -457.276552 -459.442939
          Ar  1-S  -507.249273  -521.222881  -524.342962 -526.772151
 
 
 
                                                           SCF   *
          Atom state     DH       6-311G        MC         limit
          H   2-S    -.498189     -.499810      --        -0.5
          He  1-S      --        -2.859895      --        -2.861680
          Li  2-S   -7.431736    -7.432026      --        -7.432727
          Be  1-S  -14.570907   -14.571874      --       -14.573023
          B   2-P  -24.526601   -24.527020      --       -24.529061
          C   3-P  -37.685571   -37.686024      --       -37.688619
          N   4-S  -54.397260   -54.397980      --       -54.400935
          O   3-P  -74.802707   -74.802496      --       -74.809400
          F   2-P  -99.395013   -99.394158      --       -99.409353
          Ne  1-S -128.522354  -128.522553      --      -128.547104
          Na  2-S      --           --     -161.845587  -161.858917
          Mg  1-S      --           --     -199.606558  -199.614636
          Al  2-P -241.855079       --     -241.870014  -241.876699
          Si  3-P -288.829617       --     -288.847782  -288.854380
          P   4-S -340.689043       --     -340.711346  -340.718798
          S   3-P -397.468667       --     -397.498023  -397.504910
          Cl  2-P -459.435938       --     -459.473412  -459.482088
          Ar  1-S      --           --     -526.806626  -526.817528
 
          * M.W.Schmidt and K.Ruedenberg, J.Chem.Phys. 71,
            3951-3962(1979). These are ROHF in Kh energies.
 

 
               How to do RHF, ROHF, UHF, and GVB calculations
               --- -- -- ---  ----  ---  --- --- ------------
 
                      * * * General considerations * * *
 
              These four SCF wavefunctions are all based on Fock
          operator techniques, even though some GVB runs use more
          than one determinant.  Thus all of these have an intrinsic
          N**4 time dependence, because they are all driven by
          integrals in the AO basis.  This similarity makes it
          convenient to discuss them all together.  In this section
          we will use the term HF to refer generically to any of
          these four wavefunctions, including the multi-determinate
          GVB-PP functions.  $SCF is the main input group for all
          these HF wavefunctions.
 
              As will be discussed below, in GAMESS the term ROHF
          refers to high spin open shell SCF only, but other open
          shell coupling cases are possible using the GVB code.
 
              Analytic gradients are implemented for every possible
          HF type calculation possible in GAMESS, and therefore
          numerical hessians are available for each.
 
              Analytic hessian calculation is implemented for RHF,
          ROHF, and any GVB case with NPAIR=0 or NPAIR=1.  Analytic
          hessians are more accurate, and much more quickly computed
          than numerical hessians, but require additional disk
          storage to perform an integral transformation, and also
          more physical memory.
 
              The second order Moller-Plesset energy correction
          (MP2) is implemented for RHF, UHF, and ROHF, but not any
          GVB case.  Note that since calculation of the first order
          wavefunction's density matrix is not implemented, the
          properties during an MP2 run are for the underlying HF
          (zeroth order) wavefunction.
 
              Direct SCF is implemented for every possible HF type
          calculation involving only the energy or gradient (and
          thus numerical hessians).  The direct method may not be
          used with DEM convergence, or when forming MVOs.  Direct 
          SCF may be used when analytic hessians or the MP2 energy
           correction is being computed.
 

 
                          * * * direct SCF * * *
 
              Normally, HF calculations proceed by evaluating a
          large number of two electron repulsion integrals, and
          storing these on a disk.  This integral file is read back
          in during each HF iteration to form the appropriate Fock
          operators.  In a direct HF, the integrals are not stored
          on disk, but are instead reevaluated during each HF
          iteration.  Since the direct approach *always* requires
          more CPU time, the default for DIRSCF in $SCF is .FALSE.
 
              Even though direct SCF is slower, there are at least
          three reasons why you may want to consider using it.  The
          first is that it may not be possible to store all of the
          integrals on the disk drives attached to your computer.
          Secondly, the index label packing scheme used by GAMESS
          restricts the basis set size to no more than 361 if the
          integrals are stored on disk, whereas for direct HF you
          can (in principle) use up to 2047 basis functions.
          Finally, what you are really interested in is reducing the
          wall clock time to obtain your answer, not the CPU time.
          Workstations have modest hardware (and sometimes software)
          I/O capabilities.  Other environments such as an IBM
          mainframe shared by many users may also have very poor
          CPU/wall clock performance for I/O bound jobs such as
          conventional HF.
 
              You can estimate the disk storage requirements for
          conventional HF using a P or PK file by the following
          formulae:
 
                    nint = 1/sigma * 1/8 * N**4
                    Mbytes = nint * x / 1024**2
 
          Here N is the total number of basis functions in your
          run, which you can learn from an EXETYP=CHECK run.  The
          1/8 accounts for permutational symmetry within the
          integrals.  Sigma accounts for the point group symmetry,
          and is difficult to estimate accurately.  Sigma cannot be
          smaller than 1, in no symmetry (C1) calculations.  For
          benzene, sigma would be almost six, since you generate 6
          C's and 6 H's by entering only 1 of each in $DATA.  For
          water sigma is not much larger than one, since most of the
          basis set is on the unique oxygen, and the C2v symmetry
          applies only to the H atoms.  The factor x is 12 bytes per
          integral for RHF, and 20 bytes per integral for ROHF, UHF,
          and GVB.  Finally, since integrals very close to zero need
          not be stored on disk, the actual power dependence is not
          as bad as N**4, and in fact in the limit of very large
          molecules can be as low as N**2.  Thus plugging in sigma=1
          should give you an upper bound to the actual disk space
          needed.  If the estimate exceeds your available disk
          storage, your only recourse is direct HF.
 

 
              What are the economics of direct HF?  Naively, if we
          assume the run takes 10 iterations to converge, we must
          spend 10 times more CPU time doing the integrals on each
          iteration.  However, we do not have to waste any CPU time
          reading blocks of integrals from disk, or in unpacking
          their indices.  We also do not have to waste any wall
          clock time waiting for a relatively slow mechanical device
          such as a disk to give us our data.
 
              There are some less obvious savings too, as first
          noted by Almlof.  First, since the density matrix is known
          while we are computing integrals, we can use the Schwarz
          inequality to avoid doing some of the integrals.  In a
          conventional SCF this inequality is used to avoid doing
          small integrals.  In a direct SCF it can be used to avoid
          doing integrals whose contribution to the Fock matrix is
          small (density times integral=small).  Secondly, we can
          form the Fock matrix by calculating only its change since
          the previous iteration.  The contributions to the change
          in the Fock matrix are equal to the change in the density
          times the integrals.  Since the change in the density goes
          to zero as the run converges, we can use the Schwarz
          screening to avoid more and more integrals as the
          calculation progresses.  The input option FDIFF in $SCF
          selects formation of the Fock operator by computing only
          its change from iteration to iteration.  The FDIFF option
          is not implemented for GVB since there are too many density
          matrices from the previous iteration to store, but is the
          default for direct RHF, ROHF, and UHF.
 
              So, in our hypothetical 10 iteration case, we do not
          spend as much as 10 times more time in integral
          evaluation.  Additionally, the run as a whole will not
          slow down by whatever factor the integral time is
          increased.  A direct run spends no additional time summing
          integrals into the Fock operators, and no additional time
          in the Fock diagonalizations.  So, generally speaking, a
          RHF run with 10-15 iterations will slow down by a factor
          of 2-4 times when run in direct mode.  The energy gradient
          time is unchanged by direct HF, and this is a large time
          compared to HF energy, so geometry optimizations will be
          slowed down even less.  This is really the converse of
          Amdahl's law:  if you slow down only one portion of a
          program by a large amount, the entire program slows down
          by a much smaller factor.
 
              To make this concrete, here are some times for GAMESS
          benchmark BENCH12, which is a RHF energy for a moderately
          large molecule.  The timings shown were obtained on a
          DECstation 3100 under Ultrix 3.1, which was running only
 

 
          these tests, so that the wall clock times are meaningful.
          This system is typical of Unix workstations in that it
          uses SCSI disks, and the operating system is not terribly
          good at disk I/O.  By default GAMESS stores the integrals
          on disk in the form of a P supermatrix, because this will
          save time later in the SCF cycles.  By choosing NOPK=1 in
          $INTGRL, an ordinary integral file can be used, which
          typically contains many fewer integrals, but takes more
          CPU time in the SCF.  Because the DECstation is not
          terribly good at I/O, the wall clock time for the ordinary
          integral file is actually less than when the supermatrix
          is used, even though the CPU time is longer.  The run
          takes 13 iterations to converge, the times are in seconds.
 
                                     P supermatrix   ordinary file
             # nonzero integrals      8,244,129       6,125,653
             # blocks skipped            55,841          55,841
             CPU time (ints)              709              636
             CPU time (SCF)              1289             1472
             CPU time (job total)        2123             2233
             wall time (job total)       3468             3200
 
              When the same calculation is run in direct mode
          (integrals are processed like an ordinary integral disk
          file when running direct),
 
                iteration 1:         FDIFF=.TRUE.   FDIFF=.FALSE.
             # nonzero integrals       6,117,416      6,117,416
             # blocks skipped             60,208         60,208
                iteration 13:
             # nonzero integrals       3,709,733      6,122,912
             # blocks skipped            105,278         59,415
             CPU time (job total)         6719            7851
             wall time (job total)        6764            7886
 
              Note that elimination of the disk I/O dramatically
          increases the CPU/wall efficiency.  Here's the bottom line
          on direct HF:
 
                best direct CPU / best disk CPU = 6719/2123 = 3.2
                best direct wall/ best disk wall= 6764/3200 = 2.1
 
          Direct SCF is slower than conventional SCF, but not
          outrageously so!  From the data in the tables, we can see
          that the best direct method spends about 6719-1472 = 5247
          seconds doing integrals.  This is an increase of about
          5247/636 = 8.2 in the time spent doing integrals, in a run
          which does 13 iterations (13 times evaluating integrals).
          8.2 is less than 13 because the run avoids all CPU charges 
          related to I/O, and makes efficient use of the Schwarz 
          inequality to avoid doing many of the integrals in its 
          final iterations.
 

 
                 * * * SCF convergence accelerators * * *
 
              Generally speaking, the simpler the HF function, the
          better its convergence.  In our experience, the majority
          of RHF, ROHF, and UHF runs will converge readily from
          GUESS=HUCKEL.  GVB runs typically require GUESS=MOREAD,
          although the Huckel guess occasionally works.  RHF
          convergence is the best, closely followed by ROHF.  In the
          current implementation in GAMESS, ROHF is somewhat better
          convergent than the closely related unrestricted high spin
          UHF.  For example, the radical cation of diphosphine (test
          job BENCH10) converges in 12 iterations for ROHF and 15
          iterations for UHF, starting from the neutral's closed
          shell orbitals.  GVB calculations require much more care,
          and cases with NPAIR greater than one are particularly
          difficult.
 
              Unfortunately, not all HF runs converge readily.  The
          best way to improve your convergence is to provide better
          starting orbitals!  In many cases, this means to MOREAD
          orbitals from some simpler HF case.  For example, if you
          want to do a doublet ROHF, and the HUCKEL guess does not
          seem to converge, do this:  Do an RHF on the +1 cation.
          RHF is typically more stable than ROHF, UHF, or GVB, and
          cations are usually readily convergent.  Then MOREAD the
          cation's orbitals into the neutral calculation which you
          wanted to do at first.
 
              GUESS=HUCKEL does not always guess the correct
          electronic configuration.  It may be useful to use PRTMO
          in $GUESS during a CHECK run to examine the starting
          orbitals, and then reorder them with NORDER if that seems
          appropriate.
 
              Of course, by default GAMESS uses the convergence
          procedures which are usually most effective.  Still, there
          are cases which are difficult, so the $SCF group permits
          you to select several alternative methods for improving
          convergence.  Briefly, these are
 
              EXTRAP.  This extrapolates the three previous Fock
          matrices, in an attempt to jump ahead a bit faster.  This
          is the most powerful of the old-fashioned accelerators,
          and normally should be used at the beginning of any SCF
          run.  When an extrapolation occurs, the counter at the
          left of the SCF printout is set to zero.
 
              DAMP.  This damps the oscillations between several
          successive Fock matrices.  It may help when the energy is
          seen to oscillate wildly.  Thinking about which orbitals
          should be occupied initially may be an even better way to
          avoid oscillatory behaviour.
 

 
              SHIFT.  This shifts the diagonal elements of the virtual
          part of the Fock matrix up, in an attempt to uncouple the
          unoccupied orbitals from the occupied ones.  At convergence,
          this has no effect on the orbitals, just their orbital
          energies, but will produce different (and hopefully better)
          orbitals during the iterations.
 
              RSTRCT.  This limits mixing of the occupied orbitals
          with the empty ones, especially the flipping of the HOMO
          and LUMO to produce undesired electronic configurations or
          states.  This should be used with caution, as it makes it
          very easy to converge on incorrect electronic configurations,
          especially if DIIS is also used.  If you use this, be sure 
          to check your final orbital energies to see if they are 
          sensible.  A lower energy for an unoccupied orbital than 
          for one of the occupied ones is a sure sign of problems.
 
              DIIS.  Direct Inversion in the Iterative Subspace is
          a modern method, due to Pulay, using stored error and Fock
          matrices from a large number of previous iterations to
          interpolate an improved Fock matrix.  This method was
          developed to improve the convergence at the final stages
          of the SCF process, but turns out to be quite powerful at
          forcing convergence in the initial stages of SCF as well.
          By giving ETHRSH as 10.0 in $SCF, you can practically
          guarantee that DIIS will be in effect from the first
          iteration.  The default is set up to do a few iterations
          with conventional methods (extrapolation) before engaging
          DIIS.  This is because DIIS can sometimes converge to
          solutions of the SCF equations that do not have the lowest
          possible energy.  For example, the 3-A-2 small angle state
          of SiLi2 (see M.S.Gordon and M.W.Schmidt, Chem.Phys.Lett.,
          132, 294-8(1986)) will readily converge with DIIS to a
          solution with a reasonable S**2, and an energy about 25
          milliHartree above the correct answer.  A SURE SIGN OF
          TROUBLE WITH DIIS IS WHEN THE ENERGY RISES TO ITS FINAL
          VALUE.  However, if you obtain orbitals at one point on a
          PES without DIIS, the subsequent use of DIIS with MOREAD
          will probably not introduce any problems.   Because DIIS
          is quite powerful, EXTRAP, DAMP, and SHIFT are all turned
          off once DIIS begins to work.  DEM and RSTRCT will still
          be in use, however.
 
              DEM.  Direct energy minimization should be your last
          recourse.  It explores the "line" between the current
          orbitals and those generated by a conventional change in
          the orbitals, looking for the minimum energy on that line.
          DEM should always lower the energy on every iteration,
          but is very time consuming, since each of the points
          considered on the line search requires evaluation of a
 

 
          Fock operator.  DEM will be skipped once the density
          change falls below DEMCUT, as the other methods should
          then be able to affect final convergence.   While DEM is
          working, RSTRCT is held to be true, regardless of the
          input choice for RSTRCT.  Because of this, it behooves
          you to be sure that the initial guess is occupying the
          desired orbitals.   DEM is available only for RHF.  The
          implementation in GAMESS resembles that of R.Seeger and
          J.A.Pople, J.Chem.Phys. 65, 265-271(1976).   Simultaneous
          use of DEM and DIIS resembles the ADEM-DIOS method of
          H.Sellers, Chem.Phys.Lett. 180, 461-465(1991).
 
               * * * High spin open shell SCF (ROHF) * * *
 
              Open shell SCF calculations are performed in GAMESS by
          both the ROHF code and the GVB code.  Note that when the
          GVB code is executed with no pairs, the run is NOT a true
          GVB run, and should be referred to in publications and
          discussion as a ROHF calculation.
 
              The ROHF module in GAMESS can handle any number of
          open shell electrons, provided these have a high spin
          coupling.  Some commonly occurring cases are:
 
          one open shell, doublet:
               $CONTRL SCFTYP=ROHF MULT=2 $END
 
          two open shells, triplet:
               $CONTRL SCFTYP=ROHF MULT=3 $END
 
          m open shells, high spin:
               $CONTRL SCFTYP=ROHF MULT=m+1 $END
 
              John Montgomery of United Technologies is responsible
          for the current ROHF implementation in GAMESS.  The
          following discussion is due to him:
 
              The Fock matrix in the MO basis has the form
                             closed       open        virtual
                  closed      F2      |     Fb     | (Fa+Fb)/2
                           -----------------------------------
                  open        Fb      |     F1     |    Fa
                           -----------------------------------
                  virtual   (Fa+Fb)/2 |     Fa     |    F0

          where Fa and Fb are the usual alpha and beta Fock
          matrices any UHF program produces.  The Fock operators
          for the doubly, singly, and zero occupied blocks can be
          written as
                  F2 = Acc*Fa + Bcc*Fb
                  F1 = Aoo*Fa + Boo*Fb
                  F0 = Avv*Fa + Bvv*Fb
 

 
              Some choices found in the literature for these
          canonicalization coefficients are
 
                                    Acc  Bcc  Aoo  Boo  Avv  Bvv
           Guest and Saunders       1/2  1/2  1/2  1/2  1/2  1/2
           Roothaan single matrix  -1/2  3/2  1/2  1/2  3/2 -1/2
           Davidson                 1/2  1/2   1    0    1    0
           Binkley, Pople, Dobosh   1/2  1/2   1    0    0    1
           McWeeny and Diercksen    1/3  2/3  1/3  1/3  2/3  1/3
           Faegri and Manne         1/2  1/2   1    0   1/2  1/2
 
              The choice of the diagonal blocks is arbitrary, as
          ROHF is converged when the off diagonal blocks go to zero.
          The exact choice for these blocks can however have an
          effect on the convergence rate.  This choice also affects
          the MO coefficients, and orbital energies, as the
          different choices produce different canonical orbitals
          within the three subspaces.  All methods, however, will
          give identical total wavefunctions, and hence identical
          properties such as gradients and hessians.
 
              The default coupling case in GAMESS is the Roothaan
          single matrix set.  Note that pre-1988 versions of GAMESS
          produced "Davidson" orbitals.  If you would like to fool
          around with any of these other canonicalizations, the Acc,
          Aoo, Avv and Bcc, Boo, Bvv parameters can be input as the
          first three elements of ALPHA and BETA in $SCF.
 
                 * * * Other open shell SCF cases (GVB) * * *
 
              Genuine GVB-PP runs will be discussed later in this
          section.  First, we will consider how to do open shell SCF
          with the GVB part of the program.
 
              It is possible to do other open shell cases with the
          GVB code, which can handle the following cases:
 
          one open shell, doublet:
               $CONTRL SCFTYP=GVB MULT=2 $END
               $SCF    NCO=xx NSETO=1 NO(1)=1 $END
          two open shells, triplet:
               $CONTRL SCFTYP=GVB MULT=3 $END
               $SCF    NCO=xx NSETO=2 NO(1)=1,1 $END
          two open shells, singlet:
               $CONTRL SCFTYP=GVB MULT=1 $END
               $SCF    NCO=xx NSETO=2 NO(1)=1,1 $END
 
              Note that the first two cases duplicate runs which the
          ROHF module can do better.  Note that all of these cases
          are really ROHF, since the default for NPAIR in $SCF is 0.
 
              Many open shell states with degenerate open shells
          (for example, in diatomic molecules) can be treated as
          well.  There is a sample of this in the 'Input Examples'
          section of this manual.
 

 
              If you would like to do any cases other than those
          shown above, you must derive the coupling coefficients
          ALPHA and BETA, and input them with the occupancies F in
          the $SCF group.
 
              Mariusz Klobukowski of the University of Alberta has
          shown how to obtain coupling coefficients for the GVB open
          shell program for many such open shell states.  These can
          be derived from the values in Appendix A of the book "A
          General SCF Theory" by Ramon Carbo and Josep M. Riera,
          Springer-Verlag (1978).  The basic rule is
 
                 (1)      F(i) = 1/2 * omega(i)
                 (2)  ALPHA(i) =       alpha(i)
                 (3)   BETA(i) =      - beta(i),
 
          where omega, alpha, and beta are the names used by Ramon
          in his Tables.
 
              The variable NSETO should give the number of open
          shells, and NO should give the degeneracy of each open
          shell.  Thus the 5-S state of carbon would have NSETO=2,
          and NO(1)=1,3.
 
             Some specific examples, for the lowest term in each
          of the atomic P**N configurations are
 
          !   p**1   2-P state
           $CONTRL SCFTYP=GVB  MULT=2   $END
           $SCF    NCO=xx   NSETO=1  NO=3   COUPLE=.TRUE.
                F(1)=  1.0  0.16666666666667
            ALPHA(1)=  2.0  0.33333333333333  0.00000000000000
             BETA(1)= -1.0 -0.16666666666667 -0.00000000000000  $END
          !   p**2   3-P state
           $CONTRL SCFTYP=GVB  MULT=3   $END
           $SCF    NCO=xx   NSETO=1  NO=3   COUPLE=.TRUE.
                F(1)=  1.0  0.333333333333333
            ALPHA(1)=  2.0  0.66666666666667  0.16666666666667
             BETA(1)= -1.0 -0.33333333333333 -0.16666666666667  $END
          !   p**3   4-S state
           $CONTRL SCFTYP=ROHF  MULT=4   $END
          !   p**4   3-P state
           $CONTRL SCFTYP=GVB  MULT=3   $END
           $SCF    NCO=xx   NSETO=1  NO=3   COUPLE=.TRUE.
                F(1)=  1.0  0.66666666666667
            ALPHA(1)=  2.0  1.33333333333333  0.83333333333333
             BETA(1)= -1.0 -0.66666666666667 -0.50000000000000  $END
          !   p**5   2-P state
           $CONTRL SCFTYP=GVB  MULT=2   $END
           $SCF    NCO=xx   NSETO=1  NO=3   COUPLE=.TRUE.
                F(1)=  1.0  0.83333333333333
            ALPHA(1)=  2.0  1.66666666666667  1.33333333333333
             BETA(1)= -1.0 -0.83333333333333 -0.66666666666667  $END
 

 
          Coupling constants for d**N configurations are
 
          !     d**1   2-D state
           $CONTRL SCFTYP=GVB MULT=2 $END
           $SCF    NCO=xx NSETO=1 NO=5 COUPLE=.TRUE.  F(1)=1.0,0.1
                   ALPHA(1)= 2.0, 0.20, 0.00
                    BETA(1)=-1.0,-0.10, 0.00  $END
          !     d**2   average of 3-F and 3-P states
           $CONTRL SCFTYP=GVB MULT=3 $END
           $SCF    NCO=xx NSETO=1 NO=5 COUPLE=.TRUE.  F(1)=1.0,0.2
                   ALPHA(1)= 2.0, 0.40, 0.05
                    BETA(1)=-1.0,-0.20,-0.05  $END
          !     d**3   average of 4-F and 4-P states
           $CONTRL SCFTYP=GVB MULT=3 $END
           $SCF    NCO=xx NSETO=1 NO=5 COUPLE=.TRUE.  F(1)=1.0,0.3
                   ALPHA(1)= 2.0, 0.60, 0.15
                    BETA(1)=-1.0,-0.30,-0.15  $END
          !     d**4   5-D state
           $CONTRL SCFTYP=GVB MULT=5 $END
           $SCF    NCO=xx NSETO=1 NO=5 COUPLE=.TRUE.  F(1)=1.0,0.4
                   ALPHA(1)= 2.0, 0.80, 0.30
                    BETA(1)=-1.0,-0.40,-0.30 $END
          !     d**5   6-S state
           $CONTRL SCFTYP=ROHF MULT=6 $END
          !     d**6   5-D state
           $CONTRL SCFTYP=GVB MULT=5 $END
           $SCF    NCO=xx NSETO=1 NO=5 COUPLE=.TRUE.  F(1)=1.0,0.6
                   ALPHA(1)= 2.0, 1.20, 0.70
                    BETA(1)=-1.0,-0.60,-0.50 $END
          !     d**7   average of 4-F and 4-P states
           $CONTRL SCFTYP=GVB MULT=3 $END
           $SCF    NCO=xx NSETO=1 NO=5 COUPLE=.TRUE.  F(1)=1.0,0.7
                   ALPHA(1)= 2.0, 1.40, 0.95
                    BETA(1)=-1.0,-0.70,-0.55  $END
          !     d**8   average of 3-F and 3-P states
           $CONTRL SCFTYP=GVB MULT=3 $END
           $SCF    NCO=xx NSETO=1 NO=5 COUPLE=.TRUE.  F(1)=1.0,0.8
                   ALPHA(1)= 2.0, 1.60, 1.25
                    beta(1)=-1.0,-0.80,-0.65  $end
          !     d**9   2-D state
           $CONTRL SCFTYP=GVB MULT=2 $END
           $SCF    NCO=xx NSETO=1 NO=5 COUPLE=.TRUE.  F(1)=1.0,0.9
                   ALPHA(1)= 2.0, 1.80, 1.60
                    BETA(1)=-1.0,-0.90,-0.80 $END
 
          The source for these values is R.Poirier, R.Kari, and
          I.G.Csizmadia's book "Handbook of Gaussian Basis Sets",
          Elsevier, Amsterdam, 1985.
 
          Note that GAMESS can do a proper calculation on the ground
          terms for the d**2, d**3, d**7, and d**8 configurations
          only by means of state averaged MCSCF.  For d**8, use
           $CONTRL SCFTYP=MCSCF MULT=3 $END
           $DRT    GROUP=C1 FORS=.TRUE. NMCC=xx NDOC=3 NALP=2 $END
           $GUGDIA NSTATE=10 $END
           $GUGDM2 WSTATE(1)=1,1,1,1,1,1,1,0,0,0 $END
 

 
                   * * * True GVB perfect pairing runs * * *
 
              True GVB runs are obtained by choosing NPAIR nonzero.
          If you wish to have some open shell electrons in addition
          to the geminal pairs, you may add the pairs to the end of
          any of the GVB coupling cases shown above.  The GVB module
          assumes that you have reordered your MOs into the order:
          NCO double occupied orbitals, NSETO sets of open shell
          orbitals, and NPAIR sets of geminals (with NORDER=1 in the
          $GUESS group).
 
              Each geminal consists of two orbitals and contains two
          singlet coupled electrons (perfect pairing).  The first MO
          of a geminal is probably heavily occupied (such as a
          bonding MO u), and the second is probably weakly occupied
          (such as an antibonding, correlating orbital v).  If you
          have more than one pair, you must be careful that the
          initial MOs are ordered u1, v1, u2, v2..., which is -NOT-
          the same order that RHF starting orbitals will be found
          in.  Use NORDER=1 to get the correct order.
 
              These pair wavefunctions are actually a limited form
          of MCSCF.  GVB runs are much faster than MCSCF runs,
          because the natural orbital u,v form of the wavefunction
          permits a Fock operator based optimization.  However,
          convergence of the GVB run is by no means assured.  The
          same care in selecting the correlating orbitals that you
          would apply to an MCSCF run must also be used for GVB
          runs.  In particular, look at the orbital expansions when
          choosing the starting orbitals, and check them again after
          the run converges.
 
              GVB runs will be carried out entirely in orthonormal
          natural u,v form, with strong orthogonality enforced on
          the geminals.  Orthogonal orbitals will pervade your
          thinking in both initial orbital selection, and the entire
          orbital optimization phase (the CICOEF values give the
          weights of the u,v orbitals in each geminal).  However,
          once the calculation is converged, the program will
          generate and print the nonorthogonal, generalized valence
          bond orbitals.  These GVB orbitals are an entirely
          equivalent way of presenting the wavefunction, but are
          generated only after the fact.
 
              Convergence of GVB runs is by no means as certain as
          convergence of RHF, UHF, or ROHF.  This is especially true
          when extended bases are used.  You can assist convergence
          by doing a preliminary RHF or UHF calculation (take either
          the alpha orbitals, or NOs of the latter), and use these
          orbitals for GUESS=MOREAD.   Generation of MVOs during the
          prelimnary SCF can also be advantageous.
 

 
              The total number of electrons in the GVB wavefunction
          is given by the following formula:
                  NE = 2*NCO + sum 2*F(i)*NO(i) + 2*NPAIR
                                i
          The charge is obtained by subtracting the total number of
          protons given in $DATA.  The multiplicity is implicit in
          the choice of alpha and beta constants.  Note that ICHARG
          and MULT must be given correctly in $CONTRL anyway.
 
                    * * * the special case of TCSCF * * *
 
              The wavefunction with NSETO=0 and NPAIR=1 is called
          GVB-PP(1) by Goddard, two configuration SCF (TCSCF) by
          Schaefer or Davidson, and CAS-SCF with two electrons in
          two orbitals by others.  Note that this is just semantics,
          as all of these are the identical wavefunction.  It is
          also a very important wavefunction, as TCSCF is the
          minimum treatment required to describe singlet biradicals.
          This function can be obtained with SCFTYP=MCSCF, but it is
          usually much faster to use the Fock based SCFTYP=GVB.  Due
          to its importance, the TCSCF function has been singled
          out for special attention.  The DIIS method is implemented
          for the case NSETO=0 and NPAIR=1.  Analytic hessians can
          be obtained when NPAIR=1, and in this case NSETO may be
          nonzero.
 
                * * * A caution about symmetry in GVB * * *
 
              Caution!  Some exotic calculations with the GVB
          program do not permit the use of symmetry.  The symmetry
          algorithm in GAMESS was "derived assuming that the
          electronic charge density transforms according to the
          completely symmetric representation of the point group",
          Dupuis/King, JCP, 68, 3998(1978).   This may not be true
          for certain open shell cases.  Consider the following
          correct input for the singlet-delta state of NH:
           $CONTRL SCFTYP=GVB NOSYM=1 $END
           $SCF    NCO=3 NSETO=2 NO(1)=1,1 $END
          for the x**1y**1 state, or for the x**2-y**2 state,
           $CONTRL SCFTYP=GVB NOSYM=1 $END
           $SCF    NCO=3 NPAIR=1 $END
          Neither example will work, unless you enter NOSYM=1.
          Note that if you are using coupling constants which
          have been averaged over all degenerate states which
          comprise a single term, as is done in EXAM15 and EXAM16,
          the charge density is symmetric, and these runs can
          exploit symmetry.
 
             Since GAMESS cannot at present detect that the GVB
          electronic state is not totally symmetric (or averaged
          to at least have a totally symmetric density), it is
          ---your--- responsibility to recognize that NOSYM=1
          must be chosen in the input.
 

 
                    How to do CI and MCSCF calculations
                    --- -- -- -- --- ----- ------------
 
              The majority of the input to a CI calculation is in
          the $DRT group, with $INTGRL, $CIINP, $TRANS, $CISORT,
          $GUGEM, $GUGDIA, $GUGDM, $GUGDM2, $LAGRAN groups relevant,
          but hardly ever given.  The defaults for all these latter
          groups are usually correct.  Perhaps the most interesting
          variable outside the $DRT group is NSTATE in $GUGDIA to
          include excited states in the CI computation.
 
              The MCSCF part of GAMESS is based on the CI package,
          so that most of each iteration is just a CI using the
          current set of orbitals.  The final step in each MCSCF
          iteration is the generation of improved MOs by means of a
          Newton-Raphson optimization.  Thus, all of the input just
          mentioned for a CI run except $GUGDM is still relevant,
          with the $MCSCF group being of primary concern to users.
          The $CIINP group is largely irrelevant, while the $TRFDM2
          group is possibly relevant to MCSCF gradient calculations.
          Once again, the defaults in any of the groups other than
          $DRT and $MCSCF are probably appropriate for your run.
          The most interesting variable outside $DRT and $MCSCF is
          probably WSTATE in $GUGDM2.
 
              It is very helpful to execute a EXETYP=CHECK run
          before doing any MCSCF or CI run.  The CHECK run will tell
          you the total number of CSFs, the electronic state, and
          the charge and multiplicity your $DRT generated.  The
          CHECK run also lets the program feel out the memory that
          will be required to actually do the run.  Thus the CHECK
          run can potentially prevent costly mistakes, or tell you
          when a calculation is prohibitively large.
 
                 --- Specification of the wavefunction ---
 
              The configuration state functions (CSFs) to be used
          in a CI or MCSCF calculation are specified in the $DRT by
          giving a single reference CSF, and the maximum degree of
          electron excitation from that CSF.
 
              The MOs are filled in the order FZC or MCC first,
          followed by DOC, AOS, BOS, ALP, VAL, EXT, FZV (Aufbau
          principle).  AOS, BOS, and ALP are singly occupied MOs.
          The open shell singlet couplings NAOS and NBOS must give
          the same value, and their MOs alternate.  An example is
          the input NFZC=1 NDOC=2 NAOS=2 NBOS=2 NALP=1 NVAL=3 which
          gives the reference CSF
 
               FZC,DOC,DOC,AOS,BOS,AOS,BOS,ALP,VAL,VAL,VAL
 
          This is a doublet state with five unpaired electrons.  VAL
          orbitals are unoccupied only in the reference CSF, they
          will become occupied as the other CSFs are generated.
 

 
              To perform a singles and doubles CI calculation with
          any such reference, select the electron excitation level
          IEXCIT=2.  For a normal CI, you would give all of the
          virtual MOs as VALs.
 
              Another example, for a MCSCF run, might be NMCC=3
          NDOC=3 NVAL=2, which gives the reference
 
                MCC,MCC,MCC,DOC,DOC,DOC,VAL,VAL
 
          MCSCF calculations are usually of the Full Optimized
          Reaction Space (FORS) type.  Choosing FORS=.TRUE. leads
          to an excitation level of 4, as the 6 valence electrons
          have only 4 holes available for excitation.  MCSCF runs
          typically have only a small number of VAL orbitals.  If
          you choose an excitation level IEXCIT that is smaller than
          that needed to generate the FORS space, be sure to set
          FORS=.FALSE. in $MCSCF or else VERY POOR or even NO
          CONVERGENCE will result.
 
              The next example is a first or second order CI
          wavefunction, NFZC=3 NDOC=3 NVAL=2 NEXT=-1 leads to
 
                FZC,FZC,FZC,DOC,DOC,DOC,VAL,VAL,EXT,EXT,...
 
          FOCI or SOCI is chosen by selecting the appropriate flag,
          the correct excitation level is automatically generated.
          Note that the negative number for NEXT causes all
          remaining MOs to be included in the external orbital
          space.  One way of viewing FOCI and SOCI wavefunctions is
          as all singles, or all singles and doubles from the entire
          MCSCF wavefunction as a reference.  An equivalent way of
          saying this is that all CSFs with N electrons (in this
          case N=6) distributed in the valence orbitals in all ways
          (that is the FORS MCSCF wavefunction) to make the base
          wavefunction.  To this, FOCI adds all CSFs with N-1
          electrons in active and 1 electron in external orbitals.
          SOCI adds all possible CSFs with N-2 electrons in active
          orbitals and 2 in external orbitals.  SOCI is often
          prohibitively large, but is also a very accurate
          wavefunction.
 
                           --- use of symmetry ---
 
              The CI part of the program can work with D2h, and any
          of its Abelian subgroups.  However, $DRT allows the input
          of some (but not all) higher point groups.  For these
          non-Abelian groups, the program automatically assigns the
          orbitals to an irrep in the highest possible Abelian
          subgroup.  For the other non-Abelian groups, you must at
          present select GROUP=C1.  Note that when you are computing
          a Hessian matrix, many of the displaced geometries are
          asymmetric, hence you must choose C1 in $DRT (however, be
          sure to use the highest symmetry possible in $DATA!).
 

 
              As an example, consider a molecule with Cs symmetry,
          and these two reference CSFs
 
                ...MCC...DOC DOC VAL VAL
                ...MCC...DOC AOS BOS VAL
 
          Suppose that the 2nd and 3rd active MOs have symmetries a'
          and a".  Both of these generate singlet wavefunctions,
          with 4 electrons in 4 active orbitals, but the former
          constructs 1-A' CSFs, while the latter generates 1-A"
          CSFs.  On the other hand, if the 2nd and 3rd orbitals had
          the same symmetry type, the references are equivalent.
 
                       --- selection of orbitals ---
 
              The first step is to partition the orbital space into
          core, active, and external sets, in a manner which is
          sensible for your chemical problem.  This is a bit of an
          art, and the user is referred to the references quoted at
          the end of this section.  Having decided what MCSCF to
          perform, you now must consider the more pedantic problem
          of what orbitals to begin the MCSCF calculation with.
 
              You should always start an MCSCF run with orbitals
          from some other run, by means of GUESS=MOREAD.  Do not
          expect to be able to use HCORE, MINGUESS, or EXTGUESS!
          Example 6 is a poor example, converging only because
          methylene has so much symmetry, and the basis is so small.
          If you are just beginning your MCSCF problem, use orbitals
          from some appropriate converged SCF run.  Thus, a
          realistic example of beginning a MCSCF calculation is
          examples 8 and 9.  Once you get an MCSCF to converge, you
          can and should use these MCSCF MOs (which will be Schmidt
          orthogonalized) at other nearby geometries.
 
              Starting from SCF orbitals can take a little bit of
          care.  Most of the time (but not always) the orbitals you
          want to correlate will be the highest occupied orbitals in
          the SCF.  Fairly often, however, the correlating orbitals
          you wish to use will not be the lowest unoccupied virtuals
          of the SCF.  You will soon become familiar with NORDER=1
          in $GUESS, as well as the nearly mandatory GUESS=MOREAD.
          You should also explore SYMLOC in $LOCAL, as the canonical
          SCF MOs are often spread out over regions of the molecule
          other than "where the action is".
 
              Convergence of MCSCF is by no means guaranteed.  Poor
          convergence can invariably be traced back to a poor
          initial selection of orbitals.  My advice is, before you
          even start:  "Look at the orbitals.  Then look at the
          orbitals again".  Later, if you have any trouble:  "Look
          at the orbitals some more".
 

 
              Even if you don't have any trouble, look at the
          orbitals to see if they converged to what you expected,
          and have reasonable occupation numbers.  MCSCF is by no
          means the sort of "black box" that RHF is these days, so
          LOOK VERY CAREFULLY AT YOUR RESULTS.  Since orbitals are
          very hard to look at on the screen, this means print them
          out on paper, and inspect the individual MO expansion
          coefficients!  Better yet, plot them and look at their
          nodal characteristics.
 
                    --- the iterative process ---
 
              Each MCSCF iteration consists of the following steps:
          transformation of the integrals to the current MO basis,
          sorting of these integrals, generation of the Hamiltonian
          matrix, optimization of the CI coefficients by a Davidson
          method diagonalization, generation of the second order
          density matrix, and finally orbital improvment by the
          Newton-Raphson method.  On the first iteration at the
          first geometry, you will receive the normal amount of
          printout from each of these stages, while subsequent
          iterations will produce only a single line summarizing the
          iteration.
 
              Each iteration contains MICIT Newton-Raphson orbital
          improvements.  After the first such NR step, a micro-
          iteration consists solely of an integral transformation
          over the new MOs, followed immediately by a NR orbital
          improvment, reusing the old second order density matrix.
          This avoids the CI diagonalization step, so may be of some
          use in MCSCF calculations with large CSF expansions.  If
          the number of CSFs is small, using additional micro-
          iterations is a stupid idea.
 
                      --- miscellaneous hints ---
 
              Old versions of the program required that you give
          correct input data in groups like $TRANS.  Nowadays this
          is not required, usually only $DRT and $MCSCF are given.
          Old versions of the program forced you to assign each
          orbital a symmetry code number, this is now done
          automatically.
 

 
              A very common MCSCF wavefunction, the simplest
          possible function describing a singlet diradical, has 2
          electrons in 2 active MOs.  While this wavefunction can be
          obtained in a MCSCF run (using NDOC=1 NVAL=1), it can be
          obtained much faster by use of the GVB code, with one GVB
          pair.  This GVB-PP(1) wavefunction is also known in the
          literature as two configuration SCF (TCSCF).  The two
          configuration form of this GVB is equivalent to the three
          configurations obtained in this MCSCF, as optimization in
          natural form (20 and 02) causes the coefficient of the 11
          CSF to vanish.
 
              One way to improve upon the SCF orbitals as starting
          MOs is to generate modified virtual orbitals (MVOs).
          MVOs are obtained by diagonalizing the Fock operator of a
          very positive ion, within the virtual orbital space only.
          As implemented in GAMESS, MVOs can be obtained at the end
          of any RHF, ROHF, or GVB run by setting MVOQ in $SCF
          nonzero, at the cost of a single SCF cycle.  (Generating
          MVOs in a run feeding in converged orbitals is more
          expensive, as the 2 electron integrals must be
          recomputed).  Generating MVOs will never change any of
          the occupied orbitals, but will generate more valence like
          LUMOs.
 
              It is easy to generate large numbers of CSFs with the
          $DRT input (especially SOCI), which produces a bottleneck
          in the CI diagonalization step.  If the number of CSFs is
          small (say under 10,000), then the most time consuming
          step in a CI run is the integral transformation, and in a
          MCSCF run is both the integral transformation and the
          Newton-Raphson orbital improvement.  However, larger
          numbers of CSFs will require a great deal of time in the
          diagonalization step, as well as external disk storage for
          the pieces of the Hamiltonian matrix.  The largest CI
          calculation performed at NDSU had about 85,000 CSFs,
          although we have heard of a CISD with over 200,000 CSFs
          that required 8 IBM 3380 disks to hold the Hamiltonian
          matrix.  The largest MCSCF performed at NDSU had about
          20,000 CSFs, although we have heard of one with about
          35,000 CSFs.  MCSCF typically is smaller, simply because
          this kind of run must perform a CI diagonalization once
          every iteration.
 

 
                       --- MCSCF implementation ---
 
              The MCSCF code in GAMESS is adapted from the HONDO7
          program, and was written by Michel Dupuis of IBM.  The
          implementation is of the type termed "unfolded two-step
          Newton-Raphson" by Roos.  This means that the orbital and
          CI coefficient optimizations are separated.  The latter
          are obtained in a conventional CI step, and the former by
          means of a NR orbital improvment (using the augmented
          hessian matrix).  This means convergence is not quite
          second order, but that only the orbital-orbital hessian
          terms must be computed.
 
              There are three different methods for obtaining the
          two electron contributions to the Lagrangian matrix and
          the orbital hessian in GAMESS (collectively called the NR
          matrices).  These methods compute exactly the same NR
          matrices, so the overall number of iterations is the same
          for all methods.
 
              The computational problem in forming the NR matrices
          is the combination of the second order density matrix, and
          the transformed two electron integrals.  Both of these are
          4 index matrices, but the DM2 is much smaller since its
          indices run only over occupied orbitals.  The TEIs are
          much more numerous, as two of the indices run over all
          orbitals, including unoccupied ones.
 
              METHOD=TEI is driven by the two electron integrals.
          After reading all of the DM2 into memory, this method
          begins to read the TEI file, and combines these integrals
          with the available DM2 elements.  This method uses the
          least memory, reads the DM2 and TEI files only once each,
          and is usually very much slower than the other two
          methods.
 
              METHOD=DM2 is the reverse strategy, with all the
          integrals in core, and is driven by reading DM2 elements
          from disk.  Because the number of TEIs is large, the
          implementation actually divides the TEIs into four sets,
          (ij/kl), (aj/kl), (ab/kl), and (aj/bl), where i,j,k,l run
          over the internal orbitals, and a and b are external
          orbitals.  The program must have enough memory to hold all
          the TEI's belonging to a particular set in memory.  Thus,
          construction of the NR matrices requires 4 reads of the
          TEI file, and 4 reads of the DM2.  METHOD=DM2 requires
          more memory than METHOD=TEI, and has 4 times the I/O, but
          is much faster.
 

 
              METHOD=FORMULA tries to combine the merits of both
          methods.  This method constructs a Newton formula tape on
          the first MCSCF iteration at the first geometry computed.
          This formula tape contains integer pointers telling how
          TEIs should be combined with DM2 elements to form the NR
          matrices.  If all of the TEIs will not fit into memory,
          the formula tape must be sorted after it is generated.
          The formula tape can be saved for use in later runs on the
          same molecule, see FMLFIL in $MCSCF.  This can be a very
          large disk file, which may diminish your enthusiasm for
          saving it.
 
              Once the formula tape is constructed, the "passes"
          (say M in number) begin.  A pass consists of reading as
          large a block of the TEI file as will fit into the
          available memory, then reading the related portion of the
          formula tape and the entire DM2 file for information about
          how to efficiently combine the TEI's and the DM2.  When
          all of this TEI block is processed, the next block of
          TEI's is read to begin the next "pass".  An iteration thus
          performs one complete read through the TEI and formula
          files, and M reads of the DM2.  METHOD=FORMULA uses as
          much memory as the program has available (above a certain
          minimum), by adjusting the chunk size used for the TEI
          file), has an I/O requirement somewhat larger than
          METHOD=TEI, and is perhaps a bit faster than METHOD=DM2.
          It does require additional disk space for the Newton
          formula tape.
 
              Unlike METHOD=DM2 and METHOD=TEI, METHOD=FORMULA
          requires a canonical order file of transformed integrals.
          Thus, this method does not work with the current integral
          transformation, which produces only a reverse canonical
          order file.
 
              As all three methods have advantages and
          disadvantages, all three are available.  The amount of
          memory required for each method will be included in your
          output (including EXETYP=CHECK jobs) to assist you in
          selecting a method.
 

                        --- references ---
 
              There are several review articles about second order
          MCSCF methodologies.  Of these, Roos' is a nice overview
          of the subject, while the other 3 are more detailed.
 
            1. "The Multiconfiguration (MC) SCF Method"
               B.O.Roos, in "Methods in Computational Molecular
                 Physics", edited by G.H.F.Diercksen and S.Wilson
                 D.Reidel Publishing, Dordrecht, Netherlands, 1983,
                 pp 161-187.
            2. "Optimization and Characterization of a MCSCF State"
               Olsen, Yeager, Jorgensen in "Advances in Chemical
                 Physics", vol.54, edited by I.Prigogine and S.A.Rice,
                 Wiley Interscience, New York, 1983, pp 1-176.
            3. "Matrix Formulated Direct MCSCF and Multiconfiguration
                 Reference CI Methods"
               H.-J.Werner, in "Advances in Chemical Physics", vol.69,
                  edited by K.P.Lawley, Wiley Interscience, New York,
                  1987, pp 1-62.
            4. "The MCSCF Method"   R.Shepard, ibid., pp 63-200.
 
              Some papers particularly germane to the implementation
          in GAMESS are
 
            5. "General second order MCSCF theory: A Density Matrix
                 Directed Algorithm"
               B.H.Lengsfield, III, J.Chem.Phys. 73,382-390(1980).
            6. "The use of the Augmented Matrix in MCSCF Theory"
               D.R.Yarkony, Chem.Phys.Lett. 77,634-635(1981).
 
              The next set of references discuss the choice of
          orbitals for the active space.  This is a matter of some
          importance, and needs to be understood well by the GAMESS
          user.
 
            7. "The CASSCF Method and its Application in Electronic
                 Structure Calculations"
               B.O.Roos, in "Advances in Chemical Physics", vol.69,
                  edited by K.P.Lawley, Wiley Interscience, New York,
                  1987, pp 339-445.
            8. "Are Atoms Intrinsic to Molecular Electronic
                 Wavefunctions?"
               K.Ruedenberg, M.W.Schmidt, M.M.Dombek, S.T.Elbert
                 Chem.Phys. 71, 41-49, 51-64, 65-78 (1982).
 
              The final references are simply some examples of FORS
          MCSCF applications, the latter done with GAMESS.
 
            9. D.F.Feller, M.W.Schmidt, K.Ruedenberg,
                 J.Am.Chem.Soc. 104, 960-967(1982).
           10. M.W.Schmidt, P.N.Truong, M.S.Gordon,
                 J.Am.Chem.Soc. 109, 5217-5227(1987).
 

 
                  Geometry Searches and Internal Coordinates
                  -------- -------- --- -------- -----------
 
             Stationary points are places on the potential energy
          surface with a zero gradient vector (first derivative of
          the energy with respect to nuclear coordinates).  These
          include minima (whether relative or global), better known
          to chemists as reactants, products, and intermediates; as
          well as transition states (also known as saddle points).

             The two types of stationary points have a precise
          mathematical definition, depending on the curvature of the
          potential energy surface at these points.  If all of the
          eigenvalues of the hessian matrix (second derivative
          of the energy with respect to nuclear coordinates) are
          positive, the stationary point is a minimum.  If there is
          one, and only one, negative curvature, the stationary
          point is a transition state.  Points with more than one
          negative curvature do exist, but are not important in
          chemistry.  Because vibrational frequencies are basically 
          the square roots of the curvatures, a minimum has all 
          real frequencies, and a saddle point has one imaginary 
          vibrational "frequency".

             GAMESS locates minima by geometry optimization, as
          RUNTYP=OPTIMIZE, and transition states by saddle point
          searches, as RUNTYP=SADPOINT.  In many ways these are
          similar, and in fact nearly identical FORTRAN code is used
          for both.  The term "geometry search" is used here to
          describe features which are common to both procedures.
          The input to control both RUNTYPs is found in the $STATPT
          group.
 
             As will be noted in the symmetry section below, an
          OPTIMIZE run does not always find a minimum, and a
          SADPOINT run may not find a transtion state, even though
          the gradient is brought to zero.  You can prove you have
          located a minimum or saddle point only by examining the
          local curvatures of the potential energy surface.  This
          can be done by following the geometry search with a
          RUNTYP=HESSIAN job, which should be a matter of routine.
 
                    * * * Quasi-Newton Searches * * *
 
             Geometry searches are most effectively done by what is
          called a quasi-Newton-Raphson procedure.  These methods
          assume a quadratic potential surface, and require the
          exact gradient vector and an approximation to the hessian.
          It is the approximate nature of the hessian that makes the
          method "quasi".  The rate of convergence of the geometry
          search depends on how quadratic the real surface is, and
          the quality of the hessian.  The latter is something you
          have control over, and is discussed in the next section.
 

 
             GAMESS contains two different implementations of 
          quasi-Newton procedures for finding stationary points,
          namely METHOD=STANDARD and METHOD=SCHLEGEL.  For minima
          searches, the two procedures vary in the algorithm used
          to update the hessian, and in how large steps are scaled.
          For saddle point runs, the standard method treats the 
          unique direction of smallest (hopefully negative) curvature 
          in a more sophisticated way than the Schlegel algorithm.
          For both kinds of searches, the standard method seems to
          be more robust, boasting the ability (in some cases) to 
          locate transition states even when started in a region of
          all positive local curvatures.

             The "standard method" in GAMESS has largely been 
          developed by Frank Jensen of Odense University, by 
          combining several variants of the quasi-Newton method.
          First, if the local curvatures of the approximate hessian
          are correct for the given RUNTYP, a conventional Newton-
          Raphson (NR) step is computed.  If the NR step is longer
          than the trust radius, or if the hessian curvatures are
          incorrect, a Rational Function Optimization (RFO or P-RFO)
          step will be computed.  If the RFO step is also too long,
          a restricted step is computed to ensure that the step 
          will not exceed the trust radius.  There are two rather
          different ways of thinking about this, as a min-max type
          problem (the Quadratic Approximation (QA) step) or as a
          minimization problem on the image surface (the Trust
          Radius Image Minimization (TRIM) step).  The working 
          equation is actually identical in these two methods, and
          the two rather different derivations give us more than
          one way to see why it works.  The trust radius increases
          if the ratio of the actual energy change between points 
          to the predicted change is near unity, and vice versa.
 
             The STANDARD method works so well that we use it
          exclusively, and so the SCHLEGEL method will probably be 
          omitted from some future version of GAMESS.

             You should read the papers mentioned below in order to
          understand how these methods are designed to work.  The
          first 3 papers describe the RFO and TRIM/QA algorithms
          which form the "standard method" for GAMESS' geometry 
          searches.  A good summary of various search procedures is 
          given by Bell and Crighton, and in the review by Schlegel.  
 
             1. J.Baker  J.Comput.Chem. 7,385-395(1986)
             2. T.Helgaker  Chem.Phys.Lett. 182, 305-310(1991)
             3. P.Culot, G.Dive, V.H.Nguyen, J.M.Ghuysen
                Theoret.Chim.Acta  82, 189-205(1992)
             4. H.B.Schlegel  J.Comput.Chem. 3,214(1982)
             5. S.Bell, J.S.Crighton
                J.Chem.Phys. 80, 2464-2475, (1984).
             6. H.B.Schlegel  Advances in Chemical Physics (Ab Initio
                Methods in Quantum Chemistry, Part I), volume 67,
                K.P.Lawley, Ed.  Wiley, New York, 1987, pp 249-286.
 

 
                          * * * the Hessian * * *
 
             Although quasi-Newton methods require only an
          approximation to the true hessian, the choice of this
          matrix has a great affect on convergence of the geometry
          search.  
 
             There is a procedure contained within GAMESS for
          guessing a diagonal positive definite hessian matrix,
          HESS=GUESS.  If you are using Cartesian coordinates, the
          guess hessian is 1/3 times the unit matrix.  The guess is 
          more sophisticated when internal coordinates are defined, 
          as empirical rules will be used to estimate stretching 
          and bending force constants.  Other force constants are set
          to 1/4.  The diagonal guess often works well for minima, 
          but cannot possibly find transition states (because it is 
          positive definite).  Therefore, GUESS may not be selected 
          for SADPOINT runs.
 
             Two options for providing a more accurate hessian are
          HESS=READ and CALC.  For the latter, the true hessian is
          obtained by direct calculation at the initial geometry,
          and then the geometry search begins, all in one run.  The
          READ option allows you to feed in the hessian in a $HESS
          group, as obtained by a RUNTYP=HESSIAN job.  The second
          procedure is actually preferable, as you get a chance to
          see the frequencies.  Then, if the local curvatures look
          good, you can commit to the geometry search.  Be sure to
          include a $GRAD group (if the exact gradient is available)
          in the HESS=READ job so that GAMESS can take its first step 
          immediately.
 
             Note also that you can compute the hessian at a lower
          basis set and/or wavefunction level, and read it into a
          higher level geometry search.  In fact, the $HESS group 
          could be obtained at the semiempirical level.  This trick 
          works because the hessian is 3Nx3N for N atoms, no matter 
          what atomic basis is used.  The gradient from the lower 
          level is of course worthless, as the geometry search must 
          work with the exact gradient of the wavefunction and basis 
          set in current use.  Discard the $GRAD group from the 
          lower level calculation!
 
             You often get what you pay for.  HESS=GUESS is free,
          but may lead to significantly more steps in the geometry
          search.  The other two options are more expensive at the
          beginning, but may pay back by rapid convergence to the
          stationary point.
 
             The hessian update frequently improves the hessian for a 
          few steps (especially for HESS=GUESS), but then breaks down.
          The symptoms are a nice lowering of the energy or the RMS
          gradient for maybe 10 steps, followed by crazy steps.  You
          can help by putting the best coordinates into $DATA, and
          resubmitting, to make a fresh determination of the hessian.
 

 
                      * * * Coordinate Choices * * *
        
             Optimization in cartesian coordinates has a reputation
          of converging slowly.  This is largely due to the fact
          that translations and rotations are usually left in the
          problem.  Numerical problems caused by the small eigen-
          values associated with these degrees of freedom are the 
          source of this poor convergence.  The "standard method"
          in GAMESS projects the hessian matrix to eliminate these
          degrees of freedom, which should not cause a problem.
          Nonetheless, Cartesian coordinates are in general the most
          slowly convergent coordinate system.

             The use of internal coordinates (see NZVAR in $CONTRL 
          as well as $ZMAT) also eliminates the six rotational and
          translational degrees of freedom.  Also, when internal 
          coordinates are used, the GUESS hessian is able to use 
          empirical information about bond stretches and bends.
          On the other hand, there are many possible choices for the 
          internal coordinates, and some of these may lead to much 
          poorer convergence of the geometry search than others.  
          Particularly poorly chosen coordinates may not even 
          correspond to a quadratic surface, thereby ending all hope 
          that a quasi-Newton method will converge.
 
             Internal coordinates are frequently strongly coupled.
          Because of this, Jerry Boatz has called them "infernal
          coordinates"!  A very common example to illustrate this
          might be a bond length in a ring, and the angle on the
          opposite side of the ring.  Clearly, changing one changes
          the other simultaneously.  A more mathematical definition
          of "coupled" is to say that there is a large off-diagonal
          element in the hessian.  In this case convergence may be
          unsatisfactory, especially with a diagonal GUESS hessian,
          but possibly even with an accurately calculated one.  Thus,
          a "good" set of internals is one with a diagonally dominant
          hessian.
 
             One very popular set of internal coordinates is the
          usual "model builder" Z-matrix input, where for N atoms,
          one uses N-1 bond lengths, N-2 bond angles, and N-3 bond
          torsions.  The popularity of this choice is based on its
          ease of use in specifying the initial molecular geometry.
          Typically, however, it is the worst possible choice of
          internal coordinates, and in the case of rings, is not
          even as good as Cartesian coordinates.


 
             However, GAMESS does not require this particular mix 
          of the common types.  GAMESS' only requirement is that you
          use a total of 3N-6 coordinates, chosen from these 3 basic 
          types, or several more exotic possibilities.  (Of course, 
          we mean 3N-5 throughout for linear molecules).  These
          additional types of internal coordinates include linear
          bends for 3 collinear atoms, out of plane bends, and so on.
          There is no reason at all why you should place yourself in
          a straightjacket of N-1 bonds, N-2 angles, and N-3 torsions.
          If the molecule has symmetry, be sure to use internals
          which are symmetrically related.  

             For example, the most effective choice of coordinates
          for the atoms in a four membered ring is to define all
          four sides, any one of the internal angles, and a dihedral
          defining the ring pucker.  For a six membered ring, the
          best coordinates seem to be 6 sides, 3 angles, and 3
          torsions.  The angles should be every other internal
          angle, so that the molecule can "breathe" freely.  The
          torsions should be arranged so that the central bond of
          each is placed on alternating bonds of the ring, as if
          they were pi bonds in Kekule benzene.  For a five membered
          ring, we suggest all 5 sides, 2 internal angles, again
          alternating every other one, and 2 dihedrals to fill in.
          The internal angles of necessity skip two atoms where the
          ring closes.  Larger rings should generalize on the idea
          of using all sides but only alternating angles.  If there
          are fused rings, start with angles on the fused bond, and
          alternate angles as you go around from this position.
 
             Rings and more especially fused rings can be tricky.
          For these systems, especially, we suggest the Cadillac of
          internal coordinates, the "natural internal coordinates"
          of Peter Pulay.  For a description of these, see

                P.Pulay, G.Fogarosi, F.Pang, J.E.Boggs,
                    J.Am.Chem.Soc. 101, 2550-2560 (1979).
                G.Fogarasi, X.Zhou, P.W.Taylor, P.Pulay
                    J.Am.Chem.Soc. 114, 8191-8201 (1992).

          These are linear combinations of local coordinates, except
          in the case of rings.  The examples given in these two
          papers are very thorough.

             An illustration of these types of coordinates is given
          in the example job EXAM25.INP, distributed with GAMESS.
          This is a nonsense molecule, designed to show many kinds
          of functional groups.  It is defined using standard bond
          distances with a classical Z-matrix input, and the angles
          in the ring are adjusted so that the starting value of
          the unclosed OO bond is also a standard value.



             Using Cartesian coordinates is easiest, but takes a very
          large number of steps to converge.  This however, is better
          than using the classical Z-matrix internals given in $DATA,
          which is accomplished by setting NZVAR to the correct 3N-6
          value.  The geometry search changes the OO bond length to
          a very short value on the 1st step, and the SCF fails to 
          converge.  (Note that if you have used dummy atoms in the
          $DATA input, you cannot simply enter NZVAR to optimize in
          internal coordinates, instead you must give a $ZMAT which
          involves only real atoms).

             The third choice of internal coordinates is the best set
          which can be made from the simple coordinates.  It follows
          the advice given above for five membered rings, and because
          it includes the OO bond, has no trouble with crashing this
          bond.  It takes 20 steps to converge, so the trouble of 
          generating this $ZMAT is certainly worth it compared to the
          use of Cartesians.

             Natural internal coordinates are defined in the final 
          group of input.  The coordinates are set up first for the
          ring, including two linear combinations of all angles and 
          all torsions withing the ring.  After this the methyl is
          hooked to the ring as if it were a NH group, using the
          usual terminal methyl hydrogen definitions.  The H is 
          hooked to this same ring carbon as if it were a methine.
          The NH and the CH2 within the ring follow Pulay's rules
          exactly.  The amount of input is much greater than a normal
          Z-matrix.  For example, 46 internal coordinates are given,
          which are then placed in 3N-6=33 linear combinations.  Note
          that natural internals tend to be rich in bends, and short
          on torsions.

             The energy results for the three coordinate systems 
          which converge are as follows:

            NSERCH    Cart          good Z-mat        nat. int.
             0   -48.6594935049   -48.6594935049   -48.6594935049 
             1   -48.6800538676   -48.6806631261   -48.6844234867 
             2   -48.6822702585   -48.6510215698   -48.6881175652 
             3   -48.6858299354   -48.6882945647   -48.6933816338 
             4   -48.6881499412   -48.6849667085   -48.6947421032 
             5   -48.6890226067   -48.6911899936   -48.6960508492 
             6   -48.6898261650   -48.6878047907   -48.6974654893 
             7   -48.6901936624   -48.6930608185   -48.6988318245 
             8   -48.6905304889   -48.6940607117   -48.6996847663 
             9   -48.6908626791   -48.6949137185   -48.7007217630 
            10   -48.6914279465   -48.6963767038   -48.7017826506 
            11   -48.6921521142   -48.6986608776   -48.7021265394 
            12   -48.6931136707   -48.7007305310   -48.7022386031 
            13   -48.6940437619   -48.7016095285   -48.7022533741 



            14   -48.6949546487   -48.7021531692   -48.7022564052 
            15   -48.6961698826   -48.7022080183   -48.7022564793 
            16   -48.6973813002   -48.7022454522  
            17   -48.6984850655   -48.7022492840  
            18   -48.6991553826   -48.7022503853  
            19   -48.6996239136   -48.7022507037  
            20   -48.7002269303   -48.7022508393  
            21   -48.7005379631
            22   -48.7008387759
                        ...
            50   -48.7022519950 

          from which you can see that the natural internals are
          actually the best set.  The $ZMAT exhibits upward burps
          in the energy at step 2, 4, and 6, so that for the
          same number of steps, these coordinates are always at a
          higher energy than the natural internals.

             The initial hessian generated for these three columns
          contains 0, 33, and 46 force constants.  This assists
          the natural internals, but is not the major reason for
          its superior performance.  The computed hessian at the
          final geometry of this molecule, when transformed into the
          natural internal coordinates is almost diagonal.  This
          almost complete uncoupling of coordinates is what makes
          the natural internals perform so well.  The conclusion
          is of course that not all coordinate systems are equal, 
          and natural internals are the best.  As another example,
          we have run the ATCHCP molecule, which is a popular
          geometry optimization test, due to its two fused rings:

          H.B.Schlegel, Int.J.Quantum Chem., Symp. 26, 253-264(1992)
          T.H.Fischer and J.Almlof, J.Phys.Chem. 96, 9768-9774(1992)
          J.Baker, J.Comput.Chem. 14, 1085-1100(1993)

          Here we have compared the same coordinate types, using a 
          guess hessian, or a computed hessian.  The latter set of
          runs is a test of the coordinates only, as the initial 
          hessian information is identical.  The results show clearly
          the superiority of the natural internals, which like the 
          previous example, give an energy decrease on every step:

                               HESS=GUESS   HESS=READ
          Cartesians               65          41 steps
          good Z-matrix            32          23
          natural internals        24          13
 
          A final example is phosphinoazasilatrane, with three rings
          fused on a common SiN bond, in which 112 steps in Cartesian
          space became 32 steps in natural internals.  The moral is:

              "A little brain time can save a lot of CPU time."
 

 
                       * * * The Role of Symmetry * * *
 
             At the end of a succesful geometry search, you will
          have a set of coordinates where the gradient of the energy
          is zero.  However your newly discovered stationary point 
          is not necessarily a minimum or saddle point!
 
             This apparent mystery is due to the fact that the
          gradient vector transforms under the totally symmetric
          representation of the molecular point group.  As a direct
          consequence, a geometry search is point group conserving.
          (For a proof of these statements, see J.W.McIver and
          A.Komornicki, Chem.Phys.Lett., 10,303-306(1971)).  In
          simpler terms, the molecule will remain in whatever point
          group you select in $DATA, even if the true minimum is in
          some lower point group.  Since a geometry search only
          explores totally symmetric degrees of freedom, the only
          way to learn about the curvatures for all degrees of
          freedom is RUNTYP=HESSIAN.
 
             As an example, consider disilene, the silicon analog
          of ethene.  It is natural to assume that this molecule is
          planar like ethene, and an OPTIMIZE run in D2h symmetry
          will readily locate a stationary point.  However, as a
          calculation of the hessian will readily show, this
          structure is a transition state (one imaginary frequency),
          and the molecule is really trans-bent (C2h).  A careful
          worker will always characterize a stationary point as
          either a minimum, a transition state, or some higher order
          stationary point (which is not of great interest!) by
          performing a RUNTYP=HESSIAN.

             The point group conserving properties of a geometry
          search can be annoying, as in the preceeding example, or
          advantageous.  For example, assume you wish to locate the
          transition state for rotation about the double bond in
          ethene.  A little thought will soon reveal that ethene is
          D2h, the 90 degrees twisted structure is D2d, and
          structures in between are D2.  Since the saddle point is
          actually higher symmetry than the rest of the rotational
          surface, you can locate it by RUNTYP=OPTIMIZE within D2d
          symmetry.  You can readily find this stationary point with
          the diagonal guess hessian!  In fact, if you attempt to do
          a RUNTYP=SADPOINT within D2d symmetry, there will be no
          totally symmetric modes with negative curvatures, and it
          is unlikely that the geometry search will be very well
          behaved.
 
             Although a geometry search cannot lower the symmetry,
          the gain of symmetry is quite possible.  For example, if
          you initiate a water molecule optimization with a trial
          structure which has unequal bond lengths, the geometry
          search will come to a structure that is indeed C2v (to
          within OPTTOL, anyway).  However, GAMESS leaves it up to
          you to realize that a gain of symmetry has occurred.
 

 
             In general, Mother Nature usually chooses more
          symmetrical structures over less symmetrical structures.
          Therefore you are probably better served to assume the
          higher symmetry, perform the geometry search, and then
          check the stationary point's curvatures.  The alternative
          is to start with artificially lower symmetry and see if
          your system regains higher symmetry.  The problem with
          this approach is that you don't necessarily know which
          subgroup is appropriate, and you lose the great speedups
          GAMESS can obtain from proper use of symmetry.  It is good
          to note here that "lower symmetry" does not mean simply
          changing the name of the point group and entering more
          atoms in $DATA, instead the nuclear coordinates themselves
          must actually be of lower symmetry.
 
                         * * * Practical Matters * * *
 
             Geometry searches do not bring the gradient exactly to
          zero.  Instead they stop when the largest component of the
          gradient is below the value of OPTTOL, which defaults to
          a reasonable 0.0001.   Analytic hessians usually have
          residual frequencies below 10 cm**-1 with this degree of
          optimization.  The sloppiest value you probably ever want
          to try is 0.0005.
 
             If a geometry search runs out of time, or exceeds
          NSTEP, it can be restarted.  For RUNTYP=OPTIMIZE, restart
          with the coordinates having the lowest total energy
          (do a string search on "FINAL").  For RUNTYP=SADPOINT, 
          restart with the coordinates having the smallest gradient
          (do a string search on "RMS", which means root mean square).
          These are not necessarily at the last geometry!

             The "restart" should actually be a normal run, that is
          you should not try to use the restart options in $CONTRL
          (which may not work anyway).  A geometry search can be
          restarted by extracting the desired coordinates for $DATA
          from the printout, and by extracting the corresponding
          $GRAD group from the PUNCH file.  If the $GRAD group is
          supplied, the program is able to save the time it would
          ordinarily take to compute the wavefunction and gradient
          at the initial point, which can be a substantial savings.
          There is no input to trigger reading of a $GRAD group: if
          found, it is read and used.  Be careful that your $GRAD
          group actually corresponds to the coordinates in $DATA, as
          GAMESS has no check for this.
 
             Sometimes when you are fairly close to the minimum, an
          OPTIMIZE run will take a first step which raises the
          energy, with subsequent steps improving the energy and
          perhaps finding the minimum.  The erratic first step is
          caused by the GUESS hessian.  It may help to limit the size
          of this wrong first step, by reducing its radius, DXMAX.
          Conversely, if you are far from the minimum, sometimes you
          can decrease the number of steps by increasing DXMAX.


 
             When using internals, the program uses an iterative 
          process to convert the internal coordinate change into 
          Cartesian space.  In some cases, a small change in the
          internals will produce a large change in Cartesians, and
          thus produce a warning message on the output.  If these
          warnings appear only in the beginning, there is probably 
          no problem, but if they persist you can probably devise
          a better set of coordinates.  You may in fact have one of
          the two problems described in the next paragraph.  In
          some cases (hopefully very few) the iterations to find
          the Cartesian displacement may not converge, producing a
          second kind of warning message.  The fix for this may 
          very well be a new set of internal coordinates as well,
          or adjustment of ITBMAT in $STATPT.
 
             There are two examples of poorly behaved internal 
          coordinates which can give serious problems.  The first
          of these is three angles around a central atom, when
          this atom becomes planar (sum of the angles nears 360).
          The other is a dihedral where three of the atoms are 
          nearly linear, causing the dihedral to flip between 0 and
          180.  Avoid these two situations if you want your geometry
          search to be convergent.

             Sometimes it is handy to constrain the geometry search
          by freezing one or more coordinates, via the IFREEZ array.
          For example, constrained optimizations may be useful while
          trying to determine what area of a potential energy surface 
          contains a saddle point.  If you try to freeze coordinates
          with an automatically generated $ZMAT, you need to know
          that the order of the coordinates defined in $DATA is
                y
                y  x r1
                y  x r2  x a3
                y  x r4  x a5  x w6
                y  x r7  x a8  x w9
          and so on, where y and x are whatever atoms and molecular
          connectivity you happen to be using.

                         * * * Saddle Points * * *
 
             Finding minima is relatively easy.  There are large
          tables of bond lengths and angles, so guessing starting
          geometries is pretty straightforward.  Very nasty cases
          may require computation of an exact hessian, but the
          location of most minima is straightforward.
 
             In contrast, finding saddle points is a black art.
          The diagonal guess hessian will never work, so you must
          provide a computed one.  The hessian should be computed at
          your best guess as to what the transition state (T.S.) 
          should be.  It is safer to do this in two steps as outlined 
          above, rather than HESS=CALC.  This lets you verify you 
          have guessed a structure with one and only one negative
          curvature.  Guessing a good trial structure is the hardest
          part of a RUNTYP=SADPOINT!


 
             The RUNTYP=HESSIAN's normal coordinate analysis is
          rigorously valid only at stationary points on the surface.
          This means the frequencies from the hessian at your trial
          geometry are untrustworthy, in particular the six "zero"
          frequencies corresponding to translational and rotational
          (T&R) degrees of freedom will usually be 300-500 cm**-1,
          and possibly imaginary.  The Sayvetz conditions on the
          printout will help you distinguish the T&R "contaminants"
          from the real vibrational modes.  If you have defined a
          $ZMAT, the PURIFY option within $STATPT will help zap out 
          these T&R contaminants).
 
             If the hessian at your assumed geometry does not have
          one and only one imaginary frequency (taking into account
          that the "zero" frequencies can sometimes be 300i!), then
          it will probably be difficult to find the saddle point.
          Instead you need to compute a hessian at a better guess
          for the initial geometry, or read about mode following
          below.
 
             If you need to restart your run, do so with the
          coordinates which have the smallest RMS gradient.  Note
          that the energy does not necessarily have to decrease in a
          SADPOINT run, in contrast to an OPTIMIZE run.  It is often
          necessary to do several restarts, involving recomputation
          of the hessian, before actually locating the saddle point.

             Assuming you do find the T.S., it is always a good
          idea to recompute the hessian at this structure.  As
          described in the discussion of symmetry, only totally
          symmetric vibrational modes are probed in a geometry
          search.  Thus it is fairly common to find that at your
          "T.S." there is a second imaginary frequency, which
          corresponds to a non-totally symmetric vibration.  This
          means you haven't found the correct T.S., and are back to
          the drawing board.  The proper procedure is to lower the
          point group symmetry by distorting along the symmetry
          breaking "extra" imaginary mode, by a reasonable amount.
          Don't be overly timid in the amount of distortion, or the
          next run will come back to the invalid structure.

             The real trick here is to find a good guess for the
          transition structure.  The closer you are, the better.  It
          is often difficult to guess these structures.  One way
          around this is to compute a linear least motion (LLM)
          path.  This connects the reactant structure to the product
          structure by linearly varying each coordinate.  If you
          generate about ten structures intermediate to reactants
          and products, and compute the energy at each point, you
          will in general find that the energy first goes up, and
          then down.  The maximum energy structure is a "good" guess
          for the true T.S. structure.  Actually, the success of
          this method depends on how curved the reaction path is.
 

 
                          * * * Mode Following * * *

             In certain circumstances, METHOD=STANDARD can walk from
          a region of all positive curvatures to a transition state.
          This is a tricky business, and should be attempted only
          after reading Baker's paper carefully.  Note that GAMESS
          retains all degrees of freedom in its hessian, and thus
          there is no reason to suppose the lowest mode is totally
          symmetric.  Of course, you should attempt to follow only
          totally symmetric modes with IFOLOW.  You can get a
          printout of the modes in internal coordinate space by a
          EXETYP=CHECK run, which will help you decide on the value
          of IFOLOW.
 
             Mode following is really not a substitute for the
          ability to intuit regions of the PES with a single local
          negative curvature.  When you start near a minimum, it
          matters a great deal which side of the minima you start
          from, as the direction of the search depends on the sign
          of the gradient.  Be sure to use your brain when you use
          IFOLOW!

             Figure 3 of the paper C.J.Tsai, K.D.Jordan, J.Phys.Chem.
          97, 11227-11237 (1993) is illuminating on the sensitivity
          of mode following to the initial geometry point.
 

 
                      MOPAC calculations within GAMESS
                      ----- ------------ ------ ------
 
              Parts of MOPAC 6.0 have been included in GAMESS so
          that the GAMESS user has access to three semiempirical
          wavefunctions:  MNDO, AM1 and PM3.  These wavefunctions
          are quantum mechanical in nature but neglect most two
          electron integrals, a deficiency that is (hopefully)
          compensated for by introduction of empirical parameters.
          The quantum mechanical nature of semiempirical theory
          makes it quite compatible with the ab initio methodology
          in GAMESS.  As a result, very little of MOPAC 6.0 actually
          is incorporated into GAMESS.  The part that did make it in
          is the code that evaluates
 
                1) the one- and two-electron integrals,
                2) the two-electron part of the Fock matrix,
                3) the cartesian energy derivatives, and
                4) the ZDO atomic charges and molecular dipole.
 
              Everything else is actually GAMESS:  coordinate input
          (including point group symmetry), the SCF convergence
          procedures, the matrix diagonalizer, the geometry
          searcher, the numerical hessian driver, and so on.  Most
          of the output will look like an ab initio output.
 
              It is extremely simple to perform one of these
          calculations.  All you need to do is specify GBASIS=MNDO,
          AM1, or PM3 in the $BASIS group.  Note that this not only
          picks a particular Slater orbital basis, it also selects a
          particular "hamiltonian", namely a particular parameter
          set.
 
              MNDO, AM1, and PM3 will not work with every option in
          GAMESS.  Currently the semiempirical wavefunctions support
          SCFTYP=RHF, UHF, and ROHF in any combination with
          RUNTYP=ENERGY, GRADIENT, OPTIMIZE, SADPOINT, HESSIAN, and
          IRC.  Note that all hessian runs are numerical finite
          differencing.  The MOPAC CI and half electron methods are
          not supported.
 
              Because the majority of the implementation is GAMESS
          rather than MOPAC you will notice a few improvments.
          Dynamic memory allocation is used, so that GAMESS uses far
          less memory for a given size of molecule.  The starting
          orbitals for SCF calculations are generated by a Huckel
          initial guess routine.  Spin restricted (high spin) ROHF
          can be performed.  Converged SCF orbitals will be labeled
          by their symmetry type.  Numerical hessians will make use
          of point group symmetry, so that only the symmetry unique
          atoms need to be displaced.  Infrared intensities will be
          calculated at the end of hessian runs.  We have not at
          present used the block diagonalizer during intermediate
          SCF iterations, so that the run time for a single geometry
 

 
          point in GAMESS is usually longer than in MOPAC.  However,
          the geometry optimizer in GAMESS can frequently optimize
          the structure in fewer steps than the procedure in MOPAC.
          Orbitals and hessians are punched out for convenient reuse
          in subsequent calculations.  Your molecular orbitals can
          be drawn with the PLTORB graphics program.
 
              To reduce CPU time, only the EXTRAP convergence
          accelerator is used by the SCF procdures.  For difficult
          cases, the DIIS, RSTRCT, and/or SHIFT options will work,
          but may add significantly to the run time.  With the
          Huckel guess, most calculations will converge acceptably
          without these special options.
 
              MOPAC parameters exist for the following elements.
          The quote means that these elements are treated as
          "sparkles" rather than as atoms with genuine basis
          functions.  For MNDO:
 
           H
          Li  *          B  C  N  O  F
          Na' *         Al Si  P  S Cl
           K' * ...  Zn  * Ge  *  * Br
          Rb' * ...   *  * Sn  *  *  I
          *   * ...  Hg  * Pb  *
 
                   For AM1:                         For PM3:
           H                               H
           *  *          B  C  N  O  F     *  Be         *  C  N  O  F
          Na' *         Al Si  P  S Cl    Na' Mg        Al Si  P  S Cl
           K' * ...  Zn  * Ge  *  * Br     K' * ...  Zn Ga Ge As Se Br
          Rb' * ...   *  * Sn  *  *  I    Rb' * ...  Cd In Sn Sb Te  I
          *   * ...  Hg  *  *  *          *   * ...  Hg Tl Pb Bi
 
              Semiempirical calculations are very fast.  One of the
          motives for the MOPAC implementation within GAMESS is to
          take advantage of this speed.  Semiempirical models can
          rapidly provide reasonable starting geometries for ab
          initio optimizations.  Semiempirical hessian matrices are
          obtained at virtually no computational cost, and may help
          dramatically with an ab initio geometry optimization.
          Simply use HESS=READ in $STATPT to use a MOPAC $HESS group
          in an ab initio run.
 
              It is important to exercise caution as semiempirical
          methods can be dead wrong!  The reasons for this are bad
          parameters (in certain chemical situations), and the
          underlying minimal basis set.  A good question to ask
          before using MNDO, AM1 or PM3 is "how well is my system
          modeled with an ab initio minimal basis sets, such as
          STO-3G?" If the answer is "not very well" there is a good
          chance that a semiempirical description is equally poor.
 

 
                              Molecular Properties
                              --------- ----------
 
          These two papers are of general interest:
          A.D.Buckingham, J.Chem.Phys. 30, 1580-1585(1959).
          D.Neumann, J.W.Moskowitz J.Chem.Phys. 49, 2056-2070(1968).
 
          All units are derived from the atomic units for distance
          and the monopole electric charge, as given below.
 
          distance               - 1 au = 5.291771E-09 cm
 
          monopole               - 1 au = 4.803242E-10 esu
                                  1 esu = sqrt(g-cm**3)/sec
 
          dipole                 - 1 au = 2.541766E-18 esu-cm
                                1 Debye = 1.0E-18 esu-cm
 
          quadrupole             - 1 au = 1.345044E-26 esu-cm**2
                           1 Buckingham = 1.0E-26 esu-cm**2
 
          octupole               - 1 au = 7.117668E-35 esu-cm**3
 
          electric potential     - 1 au = 9.076814E-02 esu/cm
 
          electric field         - 1 au = 1.715270E+07 esu/cm**2
                            1 esu/cm**2 = 1 dyne/esu
 
          electric field gradient- 1 au = 3.241390E+15 esu/cm**3
 
          The atomic unit for electron density is electron/bohr**3
          for the total density, and 1/bohr**3 for an orbital
          density.
 
          The atomic unit for spin density is excess alpha spins per
          unit volume, h/4*pi*bohr**3.  Only the expectation value
          is computed, with no constants premultiplying it.
 
          IR intensities are printed in Debye**2/amu-Angstrom**2.
          These can be converted into intensities as defined by
          Wilson, Decius, and Cross's equation 7.9.25, in km/mole,
          by multiplying by 42.255.  If you prefer 1/atm-cm**2, use
          a conversion factor of 171.65 instead.  A good reference
          for deciphering these units is A.Komornicki, R.L.Jaffe
          J.Chem.Phys. 1979, 71, 2150-2155.  A reference showing
          how IR intensities change with basis improvements at the
          HF level is Y.Yamaguchi, M.Frisch, J.Gaw, H.F.Schaefer,
          J.S.Binkley, J.Chem.Phys. 1986, 84, 2262-2278.
 

 
                               Localization tips
                               ------------ ----
 
              Three different orbital localization methods are
          implemented in GAMESS.  The energy and dipole based
          methods normally produce similar results, but see
          M.W.Schmidt, S.Yabushita, M.S.Gordon in J.Chem.Phys.,
          1984, 88, 382-389 for an interesting exception.  You can
          find references to the three methods at the beginning of
          this chapter.
 
              The method due to Edmiston and Ruedenberg works by
          maximizing the sum of the orbitals' two electron self
          repulsion integrals.  Most people who think about the
          different localization criteria end up concluding that
          this one seems superior.  The method requires the two
          electron integrals, transformed into the molecular orbital
          basis.  Because only the integrals involving the orbitals
          to be localized are needed, the integral transformation is
          actually not very time consuming.
 
              The Boys method maximizes the sum of the distances
          between the orbital centroids, that is the difference in
          the orbital dipole moments.
 
              The population method due to Pipek and Mezey maximizes
          a certain sum of gross atomic Mulliken populations.  This
          procedure will not mix sigma and pi bonds to yield
          localized banana bonds.  Hence it is rather easy to find
          cases where this method give rather different results than
          the Ruedenberg or Boys approach.
 
              GAMESS will localize orbitals for any kind of RHF, UHF,
          ROHF, or MCSCF wavefunctions.  The localizations will
          automatically restrict any rotation that would cause the
          energy of the wavefunction to be changed (the total
          wavefunction is left invariant).  As discussed below,
          localizations for GVB or CI functions are not permitted.
 
              The default is to freeze core orbitals.  The localized
          valence orbitals are scarcely changed if the core orbitals
          are included, and it is usually convenient to leave them
          out.  Therefore, the default localizations are:  RHF
          functions localize all doubly occupied valence orbitals.
          UHF functions localize all valence alpha, and then all
          valence beta orbitals.  ROHF functions localize all doubly
          occupied orbitals, and all singly occupied orbitals, but do
          not mix these two orbital spaces.  MCSCF functions localize
          all valence MCC type orbitals, and localize all active
          orbitals, but do not mix these two orbital spaces.  To
          recover invariant an MCSCF function, you must be using a
          FORS=.TRUE. wavefunction, and you must set GROUP=C1 in
          $DRT as the localized orbitals possess no symmetry.
 

 
              In general, GVB functions are invariant only to
          localizations of the NCO doubly occupied orbitals.  Any
          pairs must be written in natural form, so pair orbitals
          cannot be localized.  The open shells may be degenerate, so
          in general these should not be mixed.  If for some reason
          you feel you must localize the doubly occupied space, do a
          RUNTYP=PROP job.  Feed in the GVB orbitals, but tell the
          program it is SCFTYP=RHF, and enter a negative ICHARG so
          that GAMESS thinks all orbitals occupied in the GVB are
          occupied in this fictitous RHF.  Use NINA or NOUTA to
          localize the desired doubly occupied orbitals.  Orbital
          localization is not permitted for CI, because we cannot
          imagine why you would want to do that anyway.
 
              Boys localization of the core orbitals in molecules
          having elements from the third or higher row almost never
          succeeds.  Boys localization including the core for second
          row atoms will often work, since there is only one inner
          shell on these.  The Ruedenberg method should work for any
          element, although including core orbitals in the integral
          transformation is more expensive.
 
              The easiest way to do localization is in the run which
          determines the wavefunction, by selecting LOCAL=xxx in the
          $CONTRL group.  However, localization may be conveniently
          done at any time after determination of the wavefunction,
          by executing a RUNTYP=PROP job.  This will require only
          $CONTRL, $BASIS/$DATA, $GUESS (pick MOREAD), the converged
          $VEC, possibly $SCF or $DRT to define your wavefunction,
          and optionally some $LOCAL input.
 
              There is an option to restrict all rotations that would
          mix orbitals of different symmetries.  SYMLOC=.TRUE. yields
          only partially localized orbitals, but these still possess
          symmetry.  They are therefore very useful as starting
          orbitals for MCSCF or GVB-PP calculations.  Because they
          still have symmetry, these partially localized orbitals run
          as efficiently as the canonical orbitals.  Because it is
          much easier for a user to pick out the bonds which are to
          be correlated, a significant number of iterations can be
          saved, and convergence to false solutions is less likely.
 

 
                             IRC methods
                             --- -------
 
              The Intrinsic Reaction Coordinate (IRC) is defined as
          the minimum energy path connecting the reactants to products
          via the transition state.  In practice, the IRC is found by
          first locating the transition state for the reaction.  The
          IRC is then found in halves, going forward and backwards
          from the saddle point, down the steepest descent path in
          mass weighted Cartesian coordinates.  This is accomplished
          by numerical integration of the IRC equations, by a variety
          of methods to be described below.
 
              The IRC is becoming an important part of polyatomic
          dynamics research, as it is hoped that only knowledge of the
          PES in the vicinity of the IRC is needed for prediction of
          reaction rates, at least at threshhold energies.  The IRC
          has a number of uses for electronic structure purposes as
          well.  These include the proof that a certain transition
          structure does indeed connect a particular set of reactants
          and products, as the structure and imaginary frequency
          normal mode at the saddle point do not always unambiguously
          identify the reactants and products.  The study of the
          electronic and geometric structure along the IRC is also of
          interest.  For example, one can obtain localized orbitals
          along the path to determine when bonds break or form.
 
              The accuracy to which the IRC is determined is dictated
          by the use one intends for it.  Dynamical calculations
          require a very accurate determination of the path, as
          derivative information (second derivatives of the PES at
          various IRC points, and path curvature) is required later.
          Thus, a sophisticated integration method (such as AMPC4 or
          RK4), and small step sizes (STRIDE=0.05, 0.01, or even
          smaller) may be needed.  In addition to this, care should
          be taken to locate the transition state carefully (perhaps
          decreasing OPTTOL by a factor of 10), and in the initiation
          of the IRC run.  The latter might require a hessian matrix
          obtained by double differencing, certainly the hessian
          should be PURIFY'd.  Note also that EVIB must be chosen
          carefully, as decribed below.
 
              On the other hand, identification of reactants and
          products allows for much larger step sizes, and cruder
          integration methods.  In this type of IRC one might want to
          be careful in leaving the saddle point (perhaps STRIDE
          should be reduced to 0.10 or 0.05 for the first few steps
          away from the transition state), but once a few points have
          been taken, larger step sizes can be employed.  In general,
          the defaults in the $IRC group are set up for this latter,
          cruder quality IRC.  The STRIDE value for the GS2 method
          can usually be safely larger than for other methods, no 
          matter what your interest in accuracy is. 
 

 
               The simplest method of determining an IRC is linear
          gradient following, PACE=LINEAR.  This method is also known
          as Euler's method.  If you are employing PACE=LINEAR, you
          can select "stabilization" of the reaction path by the
          Ishida, Morokuma, Komornicki method.  This type of corrector
          has no apparent mathematical basis, but works rather well
          since the bisector usually intersects the reaction path at
          right angles (for small step sizes).  The ELBOW variable
          allows for a method intermediate to LINEAR and stabilized
          LINEAR, in that the stabilization will be skipped if the
          gradients at the original IRC point, and at the result of a
          linear prediction step form an angle greater than ELBOW.
          Set ELBOW=180 to always perform the stabilization.
 
               A closely related method is PACE=QUAD, which fits a
          quadratic polynomial to the gradient at the current and
          immediately previous IRC point to predict the next point.
          This pace has the same computational requirement as LINEAR,
          and is slightly more accurate due to the reuse of the old
          gradient.  However, stabilization is not possible for this
          pace, thus a stabilized LINEAR path is usually more accurate
          than QUAD.
 
              Two rather more sophisticated methods for integrating
          the IRC equations are the fourth order Adams-Moulton
          predictor-corrector (PACE=AMPC4) and fourth order Runge-
          Kutta (PACE=RK4).  AMPC4 takes a step towards the next IRC
          point (prediction), and based on the gradient found at this
          point (in the near vincinity of the next IRC point) obtains
          a modified step to the desired IRC point (correction).
          AMPC4 uses variable step sizes, based on the input STRIDE.
          RK4 takes several steps part way toward the next IRC point,
          and uses the gradient at these points to predict the next
          IRC point.  RK4 is the most accurate integration method
          implemented in GAMESS, and is also the most time consuming.

              The Gonzalez-Schlegel 2nd order method finds the next
          IRC point by a constrained optimization on the surface of
          a hypersphere, centered at 1/2 STRIDE along the gradient
          vector leading from the previous IRC point.  By construction,
          the reaction path between two successive IRC points is
          thus a circle tangent to the two gradient vectors.  The
          algorithm is much more robust for large steps than the other
          methods, so it has been chosen as the default method.  Thus,
          the default for STRIDE is too large for the other methods.  
          The number of energy and gradients need to find the next 
          point varies with the difficulty of the constrained 
          optimization, but is normally not very many points.  Be sure 
          to provide the updated hessian from the previous run when 
          restarting PACE=GS2.
          


              The number of wavefunction evaluations, and energy
          gradients needed to jump from one point on the IRC to the next
          point are summarized in the following table:
 
               PACE      # energies   # gradients
               ----      ----------   -----------
              LINEAR        1             1
          stabilized
              LINEAR        3             2
              QUAD          1             1  (+ reuse of historical
                                                      gradient)
              AMPC4         2             2  (see note)
              RK4           4             4
              GS2          2-4           2-4 (equal numbers)
 
          Note that the AMPC4 method sometimes does more than one
          correction step, with each such corection adding one more
          energy and gradient to the calculation.  You get what you
          pay for in IRC calculations:  the more energies and
          gradients which are used, the more accurate the path found.

              A description of these methods, as well as some others
          that were found to be not as good is geven by Kim Baldridge
          and Lisa Pederson, Pi Mu Epsilon Journal, 9, 513-521 (1993).
 

                                   * * * 

              All methods are initiated by jumping from the saddle
          point, parallel to the normal mode (CMODE) which has an
          imaginary frequency.  The jump taken is designed to lower
          the energy by an amount EVIB.  The actual distance taken is
          thus a function of the imaginary frequency, as a smaller
          FREQ will produce a larger initial jump.  You can simply
          provide a $HESS group instead of CMODE and FREQ, which
          involves less typing.  To find out the actual step taken for
          a given EVIB, use EXETYP=CHECK.  The direction of the jump
          (towards reactants or products) is governed by FORWRD.  Note
          that if you have decided to use small step sizes, you must
          employ a smaller EVIB to ensure a small first step.  The
          GS2 method begins by following the normal mode by one half
          of STRIDE, and then performing a hypersphere minimization
          about that point, so EVIB is irrelevant to this PACE.
 
              The only method which proves that a properly converged
          IRC has been obtained is to regenerate the IRC with a
          smaller step size, and check that the IRC is unchanged.
          Again, note that the care with which an IRC must be obtained
          is highly dependent on what use it is intended for.
 
              A small program which converts the IRC results punched
          to file IRCDATA into a format close to that required by the
          POLYRATE VTST dynamics program written in Don Truhlar's
          group is available.  For a copy of this IRCED program,
          contact Mike Schmidt.  The POLYRATE program must be obtained
          from the Truhlar group.


 
              Some key IRC references are:
          K.Ishida, K.Morokuma, A.Komornicki
                J.Chem.Phys.  66, 2153-2156 (1977)
          K.Muller
                Angew.Chem., Int.Ed.Engl.19, 1-13 (1980)
          M.W.Schmidt, M.S.Gordon, M.Dupuis
                J.Am.Chem.Soc.  107, 2585-2589 (1985)
          B.C.Garrett, M.J.Redmon, R.Steckler, D.G.Truhlar,
          K.K.Baldridge, D.Bartol, M.W.Schmidt, M.S.Gordon
                J.Phys.Chem.  92, 1476-1488(1988)
          K.K.Baldridge, M.S.Gordon, R.Steckler, D.G.Truhlar
                J.Phys.Chem.  93, 5107-5119(1989)
          C.Gonzales, H.B.Schlegel
                J.Phys.Chem.  94, 5523-5527(1990)
                J.Chem.Phys.  95, 5853-5860(1991)
                
                
 
 
              The IRC discussion closes with some practical tips:
 
              The $IRC group has a confusing array of variables, but
          fortunately very little thought need be given to most of
          them.  An IRC run is restarted by moving the coordinates of
          the next predicted IRC point into $DATA, and inserting the
          new $IRC group into your input file.  You must select the
          desired value for NPOINT.  Thus, only the first job which
          initiates the IRC requires much thought about $IRC.
 
              The symmetry specified in the $DATA deck should be the
          symmetry of the reaction path.  If a saddle point happens
          to have higher symmetry, use only the lower symmetry in
          the $DATA deck when initiating the IRC.  The reaction path
          will have a lower symmetry than the saddle point whenever
          the normal mode with imaginary frequency is not totally
          symmetric.  Be careful that the order and orientation of the
          atoms corresponds to that used in the run which generated
          the hessian matrix.
 
              If you wish to follow an IRC for a different isotope,
          use the $MASS group.  If you wish to follow the IRC in
          regular Cartesian coordinates, just enter unit masses for
          each atom.  Note that CMODE and FREQ are a function of the
          atomic masses, so either regenerate FREQ and CMODE, or
          more simply, provide the correct $HESS group.
 

 
                  Transition Moments and Spin-Orbit Coupling
                  ---------- ------- --- ---------- --------
 
              GAMESS can compute transition moments and oscillator
          strengths for the radiative transition between two CI
          wavefunctions.  The moments are computed using both the
          "length (dipole) form" and "velocity form".  GAMESS can
          also compute the one electron portion of the "microscopic
          Breit-Pauli spin orbit operator".  The two electron terms
          can be approximately accounted for by means of effective
          nuclear charges.
 
              The orbitals for the CI can be one common set of
          orbitals used by both CI states.  If one set of orbitals
          is used, the transition moment or spin-orbit coupling can
          be found for any type of CI wavefunction GAMESS can
          compute.
 
              Alternatively, two sets of orbitals (obtained from
          separate MCSCF orbital optimizations) can be used.  Two
          separate CIs will be carried out (in 1 run).  The two MO
          sets must share a common set of frozen core orbitals, and
          the CI -must- be of the complete active space type.  These
          restrictions are needed to leave the CI wavefunctions
          invariant under the necessary transformation to
          corresponding orbitals.  The non-orthogonal procedure
          implemented in GAMESS is a GUGA driven equivalent to the
          method of Lengsfield, et al.  Note that the FOCI and SOCI
          methods described by these workers are not available in
          GAMESS.
 
              If you would like to use separate orbitals for the
          states, use the FCORE option in $MCSCF in a SCFTYP=MCSCF
          optimization of the orbitals.  Typically you would
          optimize the ground state completely, and use these MCSCF
          orbitals in an optimization of the excited state, under
          the constraint of FCORE=.TRUE.  The core orbitals of such
          a MCSCF optimization should be declared MCC, rather than
          FZC, even though they are frozen.
 
              In the case of transition moments either one or two CI
          calculations are performed, necessarily on states of the
          same multiplicity.  Thus, only a $DRT1 is read.  A
          spin-orbit coupling run almost always does two CI
          calculations, as the states are usually of different
          multiplicities.  So, spin-orbit runs might read only
          $DRT1, but usually read both $DRT1 and $DRT2.  The first
          CI calculation, defined by $DRT1, must be for the state of
          lower multiplicity, with $DRT2 defining the state of
          higher multiplicity.
 

 
              You will probably have to lower the symmetry in $DRT1
          and $DRT2 to C1.  You may use full spatial symmetry in the
          DRT groups only if the two states happen to have the same
          total spatial symmetry.
 
              The transition moment and spin orbit coupling driver
          is a rather restricted path through GAMESS, in that
 
          1) GUESS=SKIP is forced, regardless of what you put in a
             $GUESS group.  The rest of this group is ignored.  The
             NOCC orbitals are read (as GUESS=MOREAD) in a $VEC1
             group, and possibly a $VEC2 group.  It is not possible
             to reorder the orbitals.
 
          2) $CIINP is not read.  The CI is hardwired to consist
             of DRT generation, integral transformation and sorting,
             Hamiltonian generation, and diagonalization.  This
             means $DRT1 (and maybe $DRT2), $TRANS, $CISORT,
             $GUGEM, and $GUGDIA input is read, and acted upon.
 
          3) The density matrices are not generated, and so no
             properties (other than the transition moment or the
             spin-orbit coupling) are computed.
 
          4) There is no restart capability provided.
 
          5) DRT input is given in $DRT1 and maybe $DRT2.
 
          6) IROOTS will determine the number of eigenvectors found
             in each CI calculation, except NSTATE in $GUGDIA will
             override IROOTS if it is larger.
 
          References:
 
          F.Weinhold, J.Chem.Phys. 54,1874-1881(1970)
 
          B.H.Lengsfield, III,  J.A.Jafri,  D.H.Phillips,
          C.W.Bauschlicher, Jr.  J.Chem.Phys. 74,6849-6856(1981)
 
          "Intramediate Quantum Mechanics, 3rd Ed." Hans A. Bethe,
          Roman Jackiw   Benjamin/Cummings Publishing, Menlo Park,
          CA (1986), chapters 10 and 11.
 
          For an application of the transition moment code:
          S.Koseki, M.S.Gordon  J.Mol.Spectrosc. 123, 392-404(1987)
 
          For more information on effective nuclear charges:
          S.Koseki, M.W.Schmidt, M.S.Gordon  
          J.Phys.Chem.  96, 10768-10772 (1992)
 
                                    * * *
 
          Special thanks to Bob Cave and Dave Feller for their
          assistance in performing check spin-orbit coupling runs
          with the MELDF programs.