File: theory_bse.html

package info (click to toggle)
abinit 7.8.2-2
  • links: PTS, VCS
  • area: main
  • in suites: jessie, jessie-kfreebsd
  • size: 278,292 kB
  • ctags: 19,095
  • sloc: f90: 463,759; python: 50,419; xml: 32,095; perl: 6,968; sh: 6,209; ansic: 4,705; fortran: 951; objc: 323; makefile: 43; csh: 42; pascal: 31
file content (690 lines) | stat: -rw-r--r-- 26,400 bytes parent folder | download
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
326
327
328
329
330
331
332
333
334
335
336
337
338
339
340
341
342
343
344
345
346
347
348
349
350
351
352
353
354
355
356
357
358
359
360
361
362
363
364
365
366
367
368
369
370
371
372
373
374
375
376
377
378
379
380
381
382
383
384
385
386
387
388
389
390
391
392
393
394
395
396
397
398
399
400
401
402
403
404
405
406
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
432
433
434
435
436
437
438
439
440
441
442
443
444
445
446
447
448
449
450
451
452
453
454
455
456
457
458
459
460
461
462
463
464
465
466
467
468
469
470
471
472
473
474
475
476
477
478
479
480
481
482
483
484
485
486
487
488
489
490
491
492
493
494
495
496
497
498
499
500
501
502
503
504
505
506
507
508
509
510
511
512
513
514
515
516
517
518
519
520
521
522
523
524
525
526
527
528
529
530
531
532
533
534
535
536
537
538
539
540
541
542
543
544
545
546
547
548
549
550
551
552
553
554
555
556
557
558
559
560
561
562
563
564
565
566
567
568
569
570
571
572
573
574
575
576
577
578
579
580
581
582
583
584
585
586
587
588
589
590
591
592
593
594
595
596
597
598
599
600
601
602
603
604
605
606
607
608
609
610
611
612
613
614
615
616
617
618
619
620
621
622
623
624
625
626
627
628
629
630
631
632
633
634
635
636
637
638
639
640
641
642
643
644
645
646
647
648
649
650
651
652
653
654
655
656
657
658
659
660
661
662
663
664
665
666
667
668
669
670
671
672
673
674
675
676
677
678
679
680
681
682
683
684
685
686
687
688
689
690
<!DOCTYPE HTML PUBLIC "-//W3C//DTD HTML 4.01 Transitional//EN">
<html><head><title>Notes on the the Bethe-Salpeter equation</title></head><body bgcolor="#ffffff">
<hr>
<h1>Notes on the Bethe-Salpeter equation</h1>

<hr>
<p>
These notes provide a brief introduction to the Bethe-Salpeter (BS) formalism
outlining the most important equations involved in the theory. 
The approach used to compute the BS kernel and the macroscopic dielectric in an implementation based 
on planewaves and norm-conserving pseudopotentials is also discussed.
</p>
<p>
The conventions used in the equations are explained in the <a href="theory_mbt.html#notations">MBT_notes</a>.
A much more consistent discussion of the theoretical aspects of the Bethe-Salpeter 
equation can be found in [1].

</p>
<h5>Copyright (C) 2006-2014 ABINIT group (MG,MS)
<br> This file is distributed under the terms of the GNU General Public License, see
~abinit/COPYING or <a href="http://www.gnu.org/copyleft/gpl.txt">
http://www.gnu.org/copyleft/gpl.txt </a>.

<br> For the initials of contributors, see ~abinit/doc/developers/contributors.txt .
</h5>

<script type="text/javascript" src="list_internal_links.js"> </script>

<h3><b>Content of the lesson</b></h3>

<p>
</p>
<ul>
<li><a href="theory_bse.html#optical_properties">1.</a> Optical properties and local field effects
  </li><li><a href="theory_bse.html#BS_equation">2.</a> The Bethe-Salpeter equation in a nutshell
  </li><li><a href="theory_bse.html#BS_solvers">3.</a> Solving the Bethe-Salpeter problem.
  </li><li><a href="theory_bse.html#kernel_mel">4.</a> Kernel matrix elements in reciprocal space 

</li>
</ul>



<h4><a name="optical_properties"></a></h4>

<h3><b> 1. Optical properties and local field effects.</b></h3>

<p>
Before discussing the Bethe-Salpeter problem, it is worth reminding some basic results concerning
the <i>ab initio</i> description of optical properties.
</p>
<p>
In frequency and reciprocal space, the microscopic dieletric function is related to the irreducible polarizability 
by the following relation

</p>
<p align="center">
        <img src="./theory_mbt/epsilon_vs_tchi.png">


</p>


from which the inverse dielectric function is obtained via matrix inversion.
<p>
The <i><b>macroscopic</b></i> dielectric function, <i>&#1108;<sub>M</sub></i><sup>LF</sup>(&#969;),
can be directly related to  the inverse of a single element, the first 
(<b>G</b><i><sub>1</sub></i>=0,<b>G</b><i><sub>2</sub></i>=0), of the inverse 
of the <i><b>microscopic</b></i> dielectric matrix by means of:

</p>
<p align="center">
        <img src="./theory_mbt/emacro_lfe.png">
</p>


The microscopic dielectric matrix is the one usually calculated within the RPA in 
a GW calculation. The optical absorption spectrum is simply given by the imaginary 
part of  <i>&#1108;<sub>M</sub></i><sup>LF</sup>(&#969;). Sometimes, especially
when comparing with experimental spectra,
the real part is simply called <i>&#1108;<sub>1</sub></i> and the imaginary part 

<i>&#1108;<sub>2</sub></i>.

<p>
Note that above equation differs from 

</p>
<p align="center">
        <img src="./theory_mbt/emacro_nlfe.png">
</p>


due to the so called local-field effects introduced by the presence of the crystalline 
environment. The former quantity <i>with</i> local fields (LF) is more accurate 
than the latter one with <i>no</i> local field (NLF). The reason the two equations are 
different is because in the first, the expression in the denominator is the first element 
of the inverse of the <i>whole</i> microscopic dielectric matrix. This element depends 
on the <i>entire</i> matrix and cannot simply be calculated by taking the inverse of 
the first element of the non-inverted matrix.

<p>
In the <a href="theory_mbt.html#RPA_Fourier_space">GW_notes</a>, we have discussed 
how to calculate the irreducible polarizability and thus the absorption spectrum within the random phase approximation (RPA).

It turns out, however, that the RPA dielectric function evaluated with Kohn-Sham orbitals and eigenvalues 
yields absorption spectra that are in quantitative, and sometimes even in qualitative, disagreement with experiments.
This is not surprising since the polarizability of the Kohn-Sham system is not expected to reproduce
the response of the fully interacting system. 
</p>
<p>
Important discrepancies with experiment are found even when the DFT gap is corrected by replacing the KS 
eigenvalues by quasiparticle energies calculated in the <i>GW</i> approximation.
This indicates that, as far as optical properties are concerned, there is some important physics 
that is not correctly captured by the <i>GW</i> approximation.
</p>
<p>
The fundamental reason for the failure of the RPA can be traced back to the neglection 
of the electron-hole interaction that, in the many-body language, should be included through the vertex function.
Replacing the vertex function with a local and instantaneous function is a too crude and
unrealistic approximation for the many-body polarizability.
In the next section we will discuss how to obtain an improved approximation for the vertex 
and therefore an improved approximation for the polarizability that takes into account many-body effects 
due to the electron-hole interaction

<br><br><br></p>
<hr><br>


<h4><a name="BS_equation"></a></h4>




<h3><b> 2. The Bethe-Salpeter equation in a nutshell.</b></h3>

<p>
A schematic picture of the different levels of the description of optical 
spectra available in ABINIT is given below.
</p>
<p style="text-align: center;">
        <img style="width: 742px; height: 433px;" src="./theory_bse/schematic_dft_gw_bse.png">
</p>


<p>The Bethe-Salpeter theory is formulated in terms of two-particle
propagators. These are four-point functions describing the motion
through the system of two particles at once. We will label them <i>L<sub>0</sub></i> 

and 
<i>L</i>, where the subscript zero indicates that the particles are non-interacting.

By restricting the starting and ending point of the particles, the two-point contraction of <i>L<sub>0</sub></i> 
gives the reducible independent-particle polarizability according to

</p>
<p align="center">
        <img src="./theory_bse/chi0_L0.png">

</p>


while the two-point contraction of <i>L</i> gives the reducible many-body polarizability.

<p align="center">
        <img src="./theory_bse/chi_L.png">
</p>


that should not be confused with the <i>irreducible</i> polarizability of the 
many-body system (denoted with a tilde in the formulas in the 

<a href="theory_mbt.html#RPA_Fourier_space">GW_notes</a>). 
The utility of working in terms of the reducible quantities is that the macroscopic 
dielectric function <i><b>with</b></i> local field effects is obtained directly from the 
reducible polarizability using

<!--
chi by taking the two-point contraction and the limit of vanishing momentum transfer.
-->

<p align="center">

        <img src="./theory_bse/em_hatchi.png">
</p>


<p>

Computing the reducible <i>L</i> directly, if possible, thus allows one to avoid the 
costly matrix inversion of the dielectric function that should otherwise be performed 
for each frequency.
<!--
that is introduced in order to avoid the costly matrix inversion that should be performed 
to obtain the macroscopic dielectric function from the dielectric matrix.
for each frequency, the BS equation is usually reformulated in terms of a generalized 
the BS equation is usually reformulated in terms of a generalized 
four-point polarizability, L, from which the macroscopic dielectric function is obtained directly using:
four-point polarizability, L, 
and the independent-particle polarizability
Roughly speaking, the BS equation allows one to obtain the four-point function P and therefore the polarizability by contraction.
From the polarizability one can then calculate the dielectric function from which the MACROSCOPIC dielectric
is obtained by matrix inversion.
In order to calculate the absorption spectrum, we need a two-point contraction of
the resulting polarizability, and the limit of vanishing momentum transfer to obtain the macroscopic dielectric function
<p>

TODO
-->


</p>
<p>
One can show that <i>L</i> satisfies the Dyson-like equation:

</p>
<p align="center">
        <img style="width: 473px; height: 33px;" src="./theory_bse/dyson_eq_4L.png">
</p>



that involves the Bethe-Salpeter kernel <i>K</i>:

<p align="center">
    
    <img style="width: 547px; height: 66px;" src="./theory_bse/bs_kernel.png">
    
</p>


where the bar symbol signifies that the Coulomb interaction has to be taken without its long-range Fourier
component at <b>G</b>=0:

<p align="center">
        <img style="width: 310px; height: 128px;" src="./theory_bse/barv.png">
</p>


<p>
As discussed in [1], local field effects (LF) are included in the formalism through 
the exchange term while the Coulomb term describes the screened electron-hole 
interaction that is responsible for the creation of the excitons. All interaction 
is in principle already contained in <i>W</i>, and the rewriting of the kernel 
in this way is mostly an algebraic trick which allows one to easily separate the 
contributions and calculate the optical spectrum both with and without the LF.
</p>
<p>
The Dyson-like equation for <i>L</i> becomes a matrix problem once a particular 
basis set is chosen to expand the four-point functions involved in the problem.
In the standard approach, the polarisabilities and the BS kernel are expressed 
in the so-called <i>transition</i> space (i.e. as products of single-particle orbitals) using:

</p>
<p align="center">

        <img style="width: 661px; height: 88px;" src="./theory_bse/eh_expansion1.png">
</p>


<p align="center">
        <img style="width: 645px; height: 55px;" src="./theory_bse/eh_expansion2.png">

</p>


where <i>n<sub>i</sub></i> is a shorthand notation to denote band, <b>k</b>-point and spin index.
The expansion is exact provided that the set of single-particle orbitals from a complete basis set in the Hilbert space. 

<p>
The main advantage of using the transition space for solving the problem is that the RPA
polarizability is diagonal in this representation

</p>
<p align="center">
        <img style="width: 551px; height: 67px;" src="./theory_bse/L0ehbasis.png">

</p>



thus leading to a considerable simplification in the mathematical formalism.
<p>
After some algebra (see [1] for a more complete derivation) one finds that,
in a system with an energy gap, the Dyson-like equation for <i>L</i> can be expressed 
in terms of an effective two-particle Hamiltonian, <i>H</i>, according to

</p>
<p align="center">
    
    <img src="./theory_bse/L_for_semiconductors.png">

</p>



The explicit form for <i>H</i> and <i>F</i> in the transition space is given by

<p align="center">
        <img style="width: 465px; height: 120px;" src="./theory_bse/Hexc_full.png">


</p>


<br>

<p align="center">
        <img style="width: 477px; height: 124px;" src="./theory_bse/F_matrix.png">
</p>


where valence states are indicated by the indices <i>v</i>,<i>v'</i> while <i>c</i>,<i>c'</i> are used for conduction bands. 
The <i>R</i> sub-matrix is Hermitian and is usually referred to as the resonant block 
while <i>C</i> (the so-called coupling block) is a complex symmetric sub-matrix.
Note that, due to the presence of the coupling term, the BS Hamiltonian is non-Hermitian 
although the operator possesses a real spectrum.

<p>
The inclusion of spin in the formalism would require an extensive discussion on its own. 
Here, for simplicity, we limit the analysis to the case of spin unpolarized semiconductors
(<a href="../input_variables/varbas.html#nsppol" target="kwimg">nsppol</a>=1).
In this particular case, the matrix elements of the resonant block are given by

</p>
<p align="center">
        <img style="width: 836px; height: 34px;" src="./theory_bse/reso_mel.png">
</p>


with single particle transition energies on the diagonal.

<!--
TODO
The <i>C</i> matrix is symmetric and its elements are given by:
<p>
-->

<p>


The matrix elements of v and <i>W</i> are defined as:

</p>
<p align="center">
        <img style="width: 900px; height: 65px;" src="./theory_bse/exchange_rspace.png">
</p>


<p align="center">
    <img style="width: 864px; height: 55px;" src="./theory_bse/coulomb_rspace.png">

</p>


Note, in particular, that only the static limit (&#969; = 0) of <i>W</i> is involved in the last expression.

<!--
It possible to show that, due to the particular symmetries of the BS kernel, the R matrix is Hermitian.
The resonant block of the excitonic Hamiltonian contains three different terms: Reso = T + Kc + Kx  where
T is a diagonal matrix with transition energies
Kc is the so-called Coulomb (or direct term) that requires the evaluation of matrix elements of the (static) screened interaction
Kx is the so-called exchange term that involves matrix elements of the modified bare Coulomb interaction
<p>
TODO
<p>
-->

<p>
The coupling matrix elements are usually smaller than the resonant ones.
This is especially true in bulk systems due to the different spatial behavior of conduction and valence states.
In solid state calculations, it is therefore common practice to ignore the <i>C</i> block (the
so-called Tamm-Dancoff approximation). 
Under this assumption the excitonic Hamiltonian is given by a Hermitian operator.

</p>
<p align="center">


    <img style="width: 981px; height: 264px;" src="./theory_bse/TDA.png">

</p>


The variable <a href="../input_variables/vargw.html#bs_coupling" target="kwimg">bs_coupling</a> defines
whether the coupling block should be included in the calculation.

<p>
The macroscopic dielectric function is obtained by contracting the many-body <i>L</i>

and then taking the optical limit of the <b>G</b>=<b>G'</b>=0 component along a particular direction
in <b>q</b>-space.
The final result reads:

</p>
<p align="center">


    <img style="width: 613px; height: 54px;" src="./theory_bse/exc_mdf.png">

</p>


where

<p align="center">

    <img style="width: 786px; height: 52px;" src="./theory_bse/oscillator_q0.png">


</p>


is the matrix element of the dipole operator in transition space.
<p>
By default the code calculates the macroscopic dielectric function taking the limit along 
six different directions in <b>q</b>-space (the three basis vectors of the reciprocal lattice and the three Cartesian
axis). It is possible to change the default directions using the variables
<a href="../input_variables/vargw.html#gw_nqlwl" target="kwimg">gw_nqlwl</a> and 

<a href="../input_variables/vargw.html#gw_qlwl" target="kwimg">gw_qlwl</a>. 

<br><br><br></p>
<hr><br>


<h4><a name="BS_solvers"></a></h4>

<h3><b> 3. Solving the Bethe-Salpeter problem.</b></h3>




<p>
At this point it is clear that the evaluation of the macroscopic dielectric function within
the BS formalism is a two-step process:

</p>
<ol>
<li>
Construction of the <i>H</i> matrix in the transition space.

<br>
<br>
</li><li>
Computation of the macroscopic dielectric function using the two equations reported at the end of the previous section.
</li>
</ol>


<p>
The flowchart of a typical Bethe-Salpeter run is schematically depicted in the diagram below:

</p>
<p align="center">

    <img src="./theory_bse/bse_flowchart.png">

</p>

<p>

The KSS file is represented with an ellipsis. 
The path on the left indicated with blue arrows represents the RPA calculation 
(<a href="../input_variables/vargs.html#optdriver" target="kwimg">optdriver</a>=3)
that produces the SCR file (see also the <a href="lesson_gw1.html">first lesson</a> of the GW tutorial).
Once the KSS and the SCR file are available, we can finally contruct the Hamiltonian and solve the
Bethe-Salpeter problem (the rectangle at the bottom of the flowchart).
</p>
<p>

For BS computations, it is common practice to simulate the self-energy corrections by employing the scissors operator 
whose value can be obtained either from experiments or from ab-initio calculations.
The scissors operator allows one to avoid a costly <i>GW</i> calculation that should performed for all the

<b>k</b>-points and bands included in the transition space
(the optional path on the right indicated with yellow arrows that 
corresponds to <a href="../input_variables/vargs.html#optdriver" target="kwimg">optdriver</a>=4).
</p>
<p>
The construction of the Hamiltonian matrix represents a significant portion of the overall CPU time 
due to the large number of transitions needed for an accurate description of the 
frequency-dependence of the polarizability.
On the other hand, also the calculation of the macroscopic dielectric function poses 
numerical difficulties since an expensive matrix inversion should be performed for each frequency.

</p>
<p>
The code implements different methods proposed in the literature to avoid the matrix inversion for each frequency.
The variable <a href="../input_variables/vargw.html#bs_algorithm" target="kwimg">bs_algorithm</a> is used to select 
among the different possibilities.
</p>
<p>
<a href="../input_variables/vargw.html#bs_algorithm" target="kwimg">bs_algorithm</a>=1
employs standard (Sca)LAPACK routines to obtain the spectral representation of <i>H</i> 

in terms of eigenvalues and right eigenvectors of <i>H</i>:

</p>
<p align="center"><img style="width: 337px; height: 218px;" src="./theory_bse/spectral_repr.png"></p>


Then, as discussed in [2], the inverse of [<i>H</i>-&#969;] is obtained according to 

<p align="center"><img style="width: 412px; height: 80px;" src="./theory_bse/resolvent.png"></p>


We do not elaborate this point further since the direct diagonalization is advantageous only in particular circumstances, 
as explained in the documentation.
<p>
<a href="../input_variables/vargw.html#bs_algorithm" target="kwimg">bs_algorithm</a>=2
avoids the diagonalization using an iterative algorithm that constructs a basis
set in which <i>H</i> is represented by a real symmetric tridiagonal matrix [3].
Without entering into detail, one can schematically represent the Haydock technique as an algorithmic procedure 
that transforms a dense (hermitian) matrix into a sparse (tridiagonal) one:

</p>
<p align="center"><img style="width: 739px; height: 181px;" src="./theory_bse/Haydock_sketch.png"></p>


Once the coefficient of the tridiagonal form are know, the macroscopic dielectric function is
evaluated in terms of the continued fraction: 

<p align="center"><img style="width: 380px; height: 147px;" src="./theory_bse/cfact.png"></p>


where a small complex shift, defined by <a href="../input_variables/vargw.html#zcut" target="kwimg">zcut</a>,
is used to avoid the singularities along the real axis.
The number of iteration required to converge is almost independent on the size of the problem, usually of the 
order of 100-200 iterations.
The algorithm involves simple matrix-vector multiplications that are efficiently performed with BLAS calls
and, most importantly, are easy to parallelize.
Another distinct advantage is that the Haydock method is less memory demanding than the direct diagonalization 
since only three temporary vectors are required instead of the full set of eigenvectors.

<p>Note that the original formulation of the method presented in [3]
assumes an Hermitian operator thus it can be used only for TDA
calculations.
We refer the reader to [4] for a generalization of the method to the
case in which the coupling term
cannot be neglected.
The main drawback of the method is that it does not give direct access
to the excitonic spectrum
hence it cannot be used to calculate binding energies or to plot the
excitonic wavefunctions.
Moreover, for the time being, <a href="../input_variables/vargw.html#bs_algorithm" target="kwimg">bs_algorithm</a>=2
cannot be used for calculations in which the coupling term is included.
</p>
<p>
<a href="../input_variables/vargw.html#bs_algorithm" target="kwimg">bs_algorithm</a>=3 employs the conjugate-gradient method
to calculate the lowest eigenstates of the Hamiltonian. 
At present, this method is still under testing and does not support calculations with the coupling term.


</p>
<h4><a name="kernel_mel"></a></h4>

<h3><b> 4. Kernel matrix elements in reciprocal space.</b></h3>



Our implementation employs planewaves to expand the periodic part of the Bloch states, <i>u</i>,
and the two-point function <i>W(r,r')</i> that becomes a <b>q</b>-dependent matrix in reciprocal-space.
The conventions used for the transforms are documented in <a href="theory_mbt.html#notations">this
section</a> of the <i>GW</i> notes.

<p>
The matrix elements of the exchange term are evaluated in reciprocal space using:

</p>
<p align="center"><img style="width: 654px; height: 70px;" src="./theory_bse/exchange_term.png"></p>



while the Coulomb term is calculated according to

<p align="center"><img style="width: 1025px; height: 73px;" src="./theory_bse/direct_term.png"></p>


The number of <b>G</b>-vectors in W and in the modified Coulomb interaction is specified through 
<a href="../input_variables/vargw.html#ecuteps" target="kwimg">ecuteps</a>
while the wavefuctions entering the oscilator matrix elements are expanded on a <b>G</b>-sphere of energy  
<a href="../input_variables/vargw.html#ecutwfn" target="kwimg">ecutwfn</a>.
The computation of the oscillator matrix elements is discussed in 

<a href="theory_mbt.html#oscillator_notes">this section</a> of the <i>GW</i> Notes.

<p>
The input variable <a href="../input_variables/vargw.html#bs_exchange_term" target="kwimg">bs_exchange_term</a>
can be used to disable the computation of the exchange term, this
option is mainly used for performing optical calculations without local
field effects.
The variable <a href="../input_variables/vargw.html#bs_coulomb_term" target="kwimg">bs_coulomb_term</a> governs 
the computation of the Coulomb term, the most CPU-demanding part due to the presence of the double sum over <b>G</b>-vectors.
</p>
<p>

It is also important to stress that, in the two equations above, the <b>k</b>-point index runs over the full
Brillouin zone hence the size of the Hamitonian is defined by the number of point in the full
Brillouin zone and not by the number of points in the irreducible wedge.

</p>
<h4><a name="dipole_mel"></a></h4>

<h3><b> 3 Matrix elements of the dipole operator.</b></h3>




The accurate calculation of optical properties require the correct treatment 
of the optical limit (<b>G</b>=0, <b>q</b> &#8594; 0) of the oscillator matrix elements. 
The computation of these terms deserves some  clarification, due to the presence of the fully nonlocal pseudopotential 
term in the Kohn-Sham Hamiltonian.
<!--
These terms are required both in RPA calculations as well as in the BS formalism.
the correct treatment of the Coulomb singularity in equation~\ref{eq:espilon_vs_tchi} deserves some 
clarification, especially if dielectric properties are calculated within the pseudopotential formalism
due to the presence of a fully nonlocal term in the Kohn-Sham Hamiltonian.
-->
<p>

A linear expansion up to the first order in <b>q</b> of the oscillator matrix element results in:

</p>
<p align="center"><img style="width: 655px; height: 46px;" src="./theory_bse/q0_expansion.png"></p>


where we have assumed <i>b<sub>1</sub> &#8800; b<sub>2</sub> </i>and the difference between the two wave functions at <b>k</b>-<b>q</b>  

and <b>k</b> has been neglected because it only introduces terms that are quadratic in <b>q</b>.

<p>
Unfortunately, the above expression cannot be directly used
because the matrix elements of the position operator are ill-defined when 
wavefunctions obey Born-von Karman periodic boundary conditions. 
For this reason, the first order contribution has to be evaluated using the equivalent expression [5]
</p>
<p>

</p>
<p align="center"><img style="width: 844px; height: 79px;" src="./theory_bse/dipole_mel.png"></p>


The term involving the momentum operator is efficiently evaluated in reciprocal space with linear
scaling in <a href="../input_variables/vargw.html#npwwfn" target="kwimg">npwwfn</a>
while the treatment of the nonlocal part of the pseudopotential is more involved and much more CPU-demanding.
<p>


The role played by this additional term is usually marginal in the case of GW calculations: 
the QP corrections are obtained by performing an integration in <b>q</b>-space
and only the <b>q</b> &#8594; 0 component of the inverse dieletric matrix is affected by the commutator of the non-local
part of the  pseudopotential.

</p>
<p>
For this reason it is common practice, especially during the <i>GW</i> convergence tests,
to neglect this contribution by setting <a href="../input_variables/vargw.html#inclvkb" target="kwimg">inclvkb</a>=0.
Strictly speaking, however, this approximation is justified only in the case of calculations in
bulk crystals provided that the BZ sampling is well converged.
Particular care has to be taken when performing <i>GW</i> calculations in non-periodic systems due to the 
reduced dimensionality of the BZ.

</p>
<p>

Please note that the commutator of the nonlocal part should ALWAYS be included
when studying optical properties, both at the RPA and at the BS level.
We suggest the use of <a href="../input_variables/vargw.html#inclvkb" target="kwimg">inclvkb</a>=2 that is 
faster and less memory demanding than the algorithm coded for
<a href="../input_variables/vargw.html#inclvkb" target="kwimg">inclvkb</a>=1.

</p>
<h3><p><b><a name="references">References</a>
</b></p></h3>


<ol>
<li><a name="onida_rmp">G. Onida, L. Reining, and A. Rubio, Rev. Mod. Phys. <b>48</b>, 601 (2002)
and references therein</a> </li><li><a name="albrecth">S. Albrecht, Ph.D. thesis, Ecole Polytechnique, Palaiseau (Paris), 1999</a> </li><li><a name="haydock2">R. Haydock, Comput. Phys. Comm. <b>20</b>, 11 (1980)</a> </li><li><a name="gruning">M. Gruning, A. Marini and X. Gonze, Nano Letters <b>9</b>, 2820 (2009) </a></li><li><a name="baroni1986">S. Baroni and R. Resta, Phys. Rev. B <b>33</b>, 7017 (1986) </a></li><!--
<li><a name="benedict">L. X. Benedict and E. L. Shirley, Phys. Rev. B <b>59</b>, 5441 (1999)</a> </li>
<li><a name="adler">S. L. Adler, Phys. Rev. <b>126</b>, 413 (1962)</a> </li>
<li><a name="wiser">N. Wiser, Phys. Rev. <b>129</b>, 72 (1963)</a> </li>
-->
</ol>

<script type="text/javascript" src="list_internal_links.js"> </script>

</body>
</html>