13.07.2015 Views

PhD thesis in English

PhD thesis in English

PhD thesis in English

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

UNIVERSITY OF BELGRADEFACULTY OF PHYSICSIvana B. VidanovićNUMERICAL STUDY OF QUANTUMGASES AT LOW TEMPERATURESDoctoral DissertationBelgrade, 2011


UNIVERZITET U BEOGRADUFIZIČKI FAKULTETIvana B. VidanovićNUMERIČKO PROUČAVANJE KVANTNIHGASOVA NA NISKIM TEMPERATURAMADoktorska disertacijaBeograd, 2011


Thesis advisor, Committee member:Dr. Antun BalažAssociate Research ProfessorInstitute of Physics BelgradeUniversity of BelgradeCommittee member:Dr. Aleksandar BogojevićAssociate Research ProfessorInstitute of Physics BelgradeUniversity of BelgradeCommittee member:Prof. Dr. Sunčica Elezović-HadžićAssociate ProfessorFaculty of PhysicsUniversity of BelgradeCommittee member:Prof. Dr. Milan KneževićProfessorFaculty of PhysicsUniversity of BelgradeThe doctoral dissertation of Ivana Vidanović was defended on December 23, 2011.


Dedicated to the memory of my father, Bora Vidanović.


lence for Computer Model<strong>in</strong>g of Complex Systems). Numerical simulationswere run on the AEGIS e-Infrastructure, supported <strong>in</strong> part by theEuropean Commission under EU FP7 projects PRACE-1IP, HP-SEEand EGI-InSPIRE.F<strong>in</strong>ally, I am <strong>in</strong>debted to my mother, sister and Borislav for their unwaver<strong>in</strong>gsupport that goes far beyond my graduate studies.


Rezime doktorske disertacijeNumeričko proučavanje kvantnih gasovana niskim temperaturamaKoncept Boze-Ajnštajn kondenzacije je u fizici prisutan još od 1924. god<strong>in</strong>e, kadaje prvi put uvedena Boze-Ajnštajn statistika za identične čestice celobrojnog sp<strong>in</strong>a[1, 2]. Već prva teorijska kvantno-mehanička razmatranja sistema ne<strong>in</strong>teragujućihčestica su ukazala na mogućnost postojanja ovog fenomena. Poznato je da fizičkeosob<strong>in</strong>e atomskih gasova jako zavise od temperature. Na sobnim temperaturama,osob<strong>in</strong>e ovih sistema su dobro opisane zakonima klasične statističke fizike i kvantnifenomeni prolaze neopaženo. Snižavanjem temperature talasna duž<strong>in</strong>a de Broljijevihtalasa materije raste i dolazi do preklapanja talasa koji odgovaraju različitimčesticama. U tom slučaju, kvantna statistika preuzima presudnu ulogu za fizičke osob<strong>in</strong>esistema. Česticama sa celobrojnim sp<strong>in</strong>om, bozonima, zakoni kvantne statistikedozvoljavaju naseljavanje istog jednočestičnog kvantnog stanja. Kako snižavamotemperaturu, sistem teži konfiguraciji sa m<strong>in</strong>imalnom energijom, a za bozone to jeupravo makroskopski naseljeno osnovno kvantno stanje. Na taj nač<strong>in</strong> ovo kvantnostanje postaje dom<strong>in</strong>antno, pa govorimo o makroskopskom kvantnom fenomenu irazlikujemo fazu sa makroskopskom naseljenošću osnovnog stanja (Boze-Ajnštajnkondenzat) i fazu bez ovog svojstava, koju nazivamo normalni gas. Fazni prelazizmedu ove dve faze naziva se Boze-Ajnštajn kondenzacija. I pored jake utemeljenostiteorijskog koncepta, do prve eksperimentalne realizacije ovog faznog prelaza jedošlo tek 1995. god<strong>in</strong>e, a rezultat je nagraden Nobelovom nagradom za fiziku 2001.Kao što smo ukratko objasnili, osnovni pojmovi o fenomenu Boze-Ajnštajn kondenzacijesu uvedeni korišćenjem modela ne<strong>in</strong>teragujućeg gasa. U realnim gasovimaje naravno nemoguće zanemariti <strong>in</strong>terakcije, zbog kojih oni na veoma niskim temperaturamaprelaze u tečno ili čvrsto stanje. Zbog toga je za eksperimentalno postizanjeBoze-Ajnštajn faznog prelaza neophodno koristiti slabo <strong>in</strong>teragujuće gasove,kao što su su, na primer, razredeni gasovi koji <strong>in</strong>teraguju kratkodometnom, Vander-Valsovom<strong>in</strong>terakcijom. U takvom sistemu, u kom je srednje rastojanje izmedučestica veliko zbog male gust<strong>in</strong>e, kvantni fenomeni dolaze do izražaja tek na veomaniskim temperaturama. Zbog mogućnosti efikasnog hladenja, za eksperimentalno


dobijanje kondenzata odabrani su atomski gasovi alkalnih metala (Rb, Li, Na, ...), ukojima je fenomen Boze-Ajnštajn kondenzacije i ostvaren po prvi put [3, 4] za tipičnegust<strong>in</strong>e čestica u opsegu od 10 18 m −3 do 10 21 m −3 (što je šest redova velič<strong>in</strong>e manjeod tipične gust<strong>in</strong>e vazduha) i temperature reda 100 nK. Postizanje ovako niskihtemperatura je zahtevalo razvijanje novih tehnika hladenja i čuvanja atoma, koji suu eksperimentu zarobljeni u zamci napravljenoj od specijalno podešenog spoljašnjegelektričnog ili magnetnog polja. Posle <strong>in</strong>tenzivnih dugogodišnjih napora, proizvodnjakvantnih gasova sač<strong>in</strong>jenih od raznih vrsta alkalnih atoma je danas standardniproces u brojnim laboratorijama širom sveta.Eksperimenti omogućavaju veoma detaljno testiranje fundamentalnih teorijskihkoncepata - kolektivnih ekscitacija Boze-Ajnštajn kondenzata, superfluidnosti (prisustvovorteksa kao odgovor sistema na spoljašnju rotaciju), osob<strong>in</strong>a faznog dijagrama[5]. Veoma važna i <strong>in</strong>teresantna osob<strong>in</strong>a ultrahladnih kvantnih gasova jeekstremno velika mogućnost kontrole svih relevantnih parametara sistema - brojačestica, gust<strong>in</strong>e, temperature, dimenzionalnosti (promenom oblika spoljašnje potencijalnezamke), a posebno je značajna mogućnost kontrolisanja jač<strong>in</strong>e <strong>in</strong>terakcijaizmedu atoma tehnikom koja se naziva Fešbah rezonanca [6]. Jednostavnompromenom spoljašnjeg magnetnog polja, efektivna <strong>in</strong>terakcija izmedu atoma se možemenjati u rasponu od mnogo redova velič<strong>in</strong>a, što č<strong>in</strong>i ove sisteme zaista jed<strong>in</strong>stvenim.U novijim eksperimentima, atomi su zarobljeni u periodičnim potencijalima, tzv.optičkim rešetkama [7]. Zahvaljujući tome, sada je moguće na nov nač<strong>in</strong> realizovatii proučavati sisteme koji su analogni sistemima poznatim iz fizike čvrstogstanja, a koje još uvek ne razumemo u potpunosti (npr. visoko-temperaturna superprovodljivost).Upravo zato se kaže da ultrahladni kvantni gasovi predstavljajuFajnmanove “kvantne simulatore” [8]. Intenzivan eksperimentalni razvoj i realizacijanove ultrahladne faze materije predstavljaju snažan podsticaj za nova, <strong>in</strong>terdiscipl<strong>in</strong>arnateorijska istraživanja.Osnovni cilj ove teze je podrobno razumevanje dva zanimljiva fizička scenarijaza manipulaciju hladnim bozonskim atomima, koja su predmet i nedavnih eksperimentalnihistraživanja. Prvo smo razmotrili fazni dijagram rotirajućeg idealnogbozonskog gasa u anharmonijskom potencijalu, dok se druga noseća tema teze bav<strong>in</strong>el<strong>in</strong>earnim osob<strong>in</strong>ama kolektivnih bozonskih moda, koje su pobudene harmonijskommodulacijom <strong>in</strong>terakcije.Da bismo ostvarili ove ciljeve, najpre smo u Poglavlju 2 razradili detalje numeričkogmetoda koji nam na efikasan nač<strong>in</strong> pruža <strong>in</strong>formaciju o velikom broju


svojstvenih stanja kvantnog sistema. Tačno poznavanje energetskih nivoa nam jeneophodno radi preciznog odredivanja faznog dijagrama Boze-Ajnštajn kondenzata.Metod koji smo koristili je zasnovan na egzaktnoj dijagonalizaciji evolucionog operatora[9], a da bi postigli njegovu optimalnu upotrebu, detaljno smo analiziraligreške koje nastaju pri korišćenju ovog metoda iz dva razloga: greške nastale usleduvodenja prostorne diskretizacije, kao i greške pri računanju matričnih elemenataevolucionog operatora. Jedan od naših glavnih rezultata je mnogo optimalnijeponašanje diskretizacione greške ovog metoda u odnosu na standardni metod dijagonalizacijeprostorno diskretizovanog Hamiltonijana. Detaljnim analitičkim <strong>in</strong>umeričkim razmatranjem, pokazali smo da dijagonalizacija diskretizovanog evolucionogoperatora pokazuje neperturbativno malu diskretizacionu grešku, koja opadaeksponencijalno sa 1/∆ 2 , gde je ∆ korak prostorne diskretizacije, dok standardnimetodi imaju grešku koja pol<strong>in</strong>omijalno zavisi od ∆. Ovo je osnovni razlog zbogkog je mnogo optimalnije koristiti dijagonalizaciju evolucionog operatora. Glavnateškoća u primeni ovog pristupa - precizno računanje matričnih elemanata evolucionogoperatora, tj. amplituda prelaza, direktno se razrešava primenom ranije uvedenogmetoda efektivnih dejstava [10, 11], koja nam daju razvoj amplitude prelazapo kratkom vremenu propagacije do veoma visokog nivoa. Veliku efikasnost ovogmetoda smo demonstrirali na nekoliko jednodimenzionalnih i dvodimenzionalnihmodela.U poglavlju 3 primenili smo prethodno opisani metod na ispitivanje faznogdijagrama rotirajućih bozona u nestandardnom spoljašnjem potencijalu. Naime,najčešće korišćene potencijalne zamke su harmonijskog oblika, i o Boze-Ajnštajnkondenzaciji u ovakvim zamkama se već puno zna. Rotacija kvantnog gasa je jedanod nač<strong>in</strong>a da se ostvare jako korelisane faze materije [7] i od velikog je značaja.Medutim, jedna od posledica rotacije je pojava dekonf<strong>in</strong>irajuće centrifugalne komponenteu potencijalu, koja za velike frekvencije rotacije (veće of frekvencije harmonijskepotencijalne zamke) dovodi do razletanja čestica gasa i gubitka kondenzata.Da bi se to izbeglo, u nedavnom eksperimentu [12] je upotrebljen dodatnikvartični potencijal za formiranje potencijalne zamke. U zavisnosti od frekvencijerotacije, ukupan efektivni potencijal menja oblik od konveksnog potencijala sa jednimm<strong>in</strong>imumom do potencijala koji ima oblik meksičkog šešira. Primenom egzaktnedijagonalizacije evolucionog operatora, proučavali smo kako promena spoljašnjeg potencijalautiče na temperaturu Boze-Ajnštajn kondenzacije, na raspodelu čestica uzamci i na rezultate eksperimentalnih merenja.


Uticaj slabih <strong>in</strong>terakcija na fenomen Boze-Ajnštajn kondenzacije je predmet razmatranjaPoglavlja 4. Ovo poglavlje je preglednog tipa, i u njemu smo predstaviliHartri-Fok opis bozonskog sistema. Hartri-Fok predstavlja jednu od aproksimacija uteoriji srednjeg polja, i u okviru nje smo opisali sistem ultrahladnih bozona na nultojtemperaturi, kao i u okol<strong>in</strong>i Boze-Ajnštajn faznog prelaza. Za nultu temperaturusmo izveli čuvenu Gros-Pitaevski jednač<strong>in</strong>u [13, 14] - nel<strong>in</strong>earnu parcijalnu diferencijalnujednač<strong>in</strong>u koja opisuje ponašanje makroskopse talasne funkcije kondnezata.Za konačne temperature predstavili smo nekoliko često korišćenih implementacijaaproksimacije srednjeg polja i naveli njihove prednosti i nedostatke. Aproksimativnimetodi ove vrste se veoma često koriste u <strong>in</strong>terpretaciji eksperimentalnih merenja,što im daje veliki značaj. Nedavni eksperimentalni napredak u uočavanju efekatakoji su izvan opisa teorije srednjeg polja zahteva popravke standardno korišćenihaproksimacija u najskorijoj budućnosti.Jedan od osnovnih nač<strong>in</strong>a karaketrizacije faza materije, kako eksperimentalnotako i teorijski, su osob<strong>in</strong>e njihovog ekscitaciong spektra, a posebno su <strong>in</strong>teresantnei važne kolektivne mode. U sistemima hladnih gasova, kolektivne mode se običnopobuduju modulacijom spoljašnjeg potencijala zamke, dok je u nedavnom eksperimentalnomradu [15] pobudivanje kolektivnih moda ostvareno novim pristupom- harmonijskom modulacijom <strong>in</strong>terakcije. U osnovi primenjenog eksperimentalnogmetoda je tehnika Fešbah rezonance, kojom se jač<strong>in</strong>a kratkodometne <strong>in</strong>terakcijemenja u vremenu usled modulacije spoljašnjeg magnetnog polja. Kao posledicu,imamo oscilacije velič<strong>in</strong>e bozonskog oblaka, koje su u eksperimentu merene. U zavisnostiod vrednosti spoljašnje frekvencije upotrebljene za modulaciju <strong>in</strong>terakcije,možemo da dobijemo l<strong>in</strong>earni odgovor sistema ili rezonantno ponašanje karakterisanovelikim amplitudama oscilacija. Kako je osnovna jednač<strong>in</strong>a koja opisuje d<strong>in</strong>amikuovakvog sistema nel<strong>in</strong>earna, u slučaju velikih oscilacija očekujemo izraženenel<strong>in</strong>earne efekte. U poglavlju 5 smo numerički simulirali d<strong>in</strong>amiku sistema i identifikoval<strong>in</strong>el<strong>in</strong>earne karakteristike dobijenih ekscitacionih spektara: pored osnovnihmoda, pojavljuju se viši harmonici, kao i l<strong>in</strong>earne komb<strong>in</strong>acije različitih moda, anajizraženiji nel<strong>in</strong>earni efekati su pomeraji u frekvencijama ekscitovanih moda uodnosu na vrednosti izračunate u l<strong>in</strong>earnom režimu. Za kvantitativno objašnjenjenel<strong>in</strong>earnih efekata razvili smo perturbativi pristup u kom je mali parametar amplitudamodulacije. Razvijeni perturbativni pristup je baziran na Poenkare-L<strong>in</strong>dštetmetodu, i njegovom primenom smo našli analitičke izraze za nel<strong>in</strong>earne pomerajesvojstvenih frekvencija u bliz<strong>in</strong>i rezonanci.


Ključne reči: hladni kvantni gasovi, Boze-Ajnštajn kondenzacija, efektivnodejstvo, egzaktna dijagonalizacija, teorija perturbacija, nel<strong>in</strong>earna d<strong>in</strong>amikaNaučna oblast: FizikaUža naučna oblast: Fizika kondenzovanog stanja materijeUDK broj: 538.9


Abstract of the doctoral dissertationNumerical study of quantum gasesat low temperaturesThe concept of Bose-E<strong>in</strong>ste<strong>in</strong> condensation was <strong>in</strong>troduced <strong>in</strong> 1924, at the sametime as Bose-E<strong>in</strong>ste<strong>in</strong> statistics, applicable to the <strong>in</strong>teger-sp<strong>in</strong> particles [1, 2]. Alreadyfirst theoretical quantum-mechanical considerations po<strong>in</strong>ted to the existence ofthis phenomenon. It is well known that physical properties of atomic gases stronglydepend on the temperature. At room temperatures the properties of these systemscan be described by the classical statistical physics, and quantum features are negligible.However, as the temperature decreases, the wavelengths of the de Brogliemater waves <strong>in</strong>crease, lead<strong>in</strong>g to the overlap of waves correspond<strong>in</strong>g to different particles.In this case, quantum statistics plays a dom<strong>in</strong>ant role. Accord<strong>in</strong>g to the rulesof quantum statistics, <strong>in</strong>teger-sp<strong>in</strong> particles, bosons, are allowed to occupy the sames<strong>in</strong>gle-particle quantum states. With the decrease <strong>in</strong> the temperature, the systemseeks the m<strong>in</strong>imal-energy configuration, and for bosons this is a macroscopicallyoccupied s<strong>in</strong>gle-particle ground state. Such a quantum state becomes dom<strong>in</strong>ant andthe occurrence is designated as a macroscopic quantum phenomenon. Accord<strong>in</strong>gly,we dist<strong>in</strong>guish a phase with a macroscopic occupation of the ground state (Bose-E<strong>in</strong>ste<strong>in</strong> condensate) and a phase without this feature, which is called a normal gas.The phase transition between the two phases is Bose-E<strong>in</strong>ste<strong>in</strong> condensation. In spiteof the firm theoretical foundation of the concept from the beg<strong>in</strong>n<strong>in</strong>g, first direct experimentalobservation was achieved only <strong>in</strong> 1995, and the result was recognized bythe Nobel prize for physics <strong>in</strong> 2001.As briefly expla<strong>in</strong>ed, the basic notion of Bose-E<strong>in</strong>ste<strong>in</strong> condensation was <strong>in</strong>troducedus<strong>in</strong>g a model of non<strong>in</strong>teract<strong>in</strong>g gas. Of course, <strong>in</strong> realistic gases, <strong>in</strong>teractionscan not be neglected and only due to them the gas becomes liquid or solid at lowtemperatures. In order to rema<strong>in</strong> close to the non<strong>in</strong>teract<strong>in</strong>g gas description <strong>in</strong>experiments, it is essential to use weakly <strong>in</strong>teract<strong>in</strong>g gases, such as dilute gases <strong>in</strong>teract<strong>in</strong>gvia short-range, van der Waals <strong>in</strong>teraction. In this type of systems, withlong <strong>in</strong>terparticle distances due to diluteness, the quantum phenomena become relevantonly at very low temperatures. Due to highly efficient cool<strong>in</strong>g techniques,


the gases of alkali metals (Rb, Li, Na, ...) were selected as most suitable candidatesand Bose-E<strong>in</strong>ste<strong>in</strong> condensation was observed for the first time [3, 4] <strong>in</strong> such systemswith the typical particle densities from 10 18 m −3 to 10 21 m −3 (six orders of magnitudelower than the density of air) and temperatures of the order of 100 nK. In order toreach this extreme low-temperature regime, it was necessary to develop many newcool<strong>in</strong>g and trapp<strong>in</strong>g techniques: <strong>in</strong> experiments, atoms are conf<strong>in</strong>ed us<strong>in</strong>g speciallydesigned configurations of external magnetic or electric fields. After many yearsof <strong>in</strong>tensive experimental efforts, the achievement of Bose-E<strong>in</strong>ste<strong>in</strong> condensation <strong>in</strong>cold alkali vapors is nowadays a common technique <strong>in</strong> laboratories all over the world.The experiments make possible very detailed tests of fundamental theoreticalconcepts - collective excitations of a Bose-E<strong>in</strong>ste<strong>in</strong> condensate, superfluidity (theappearance of vortices as a response to rotation), the properties of a phase diagram[5]. A very important and <strong>in</strong>terest<strong>in</strong>g feature of ultracold quantum gases is a possibilityto control relevant parameters of the system over many orders of magnitude.Basically, all parameters can be tuned: number of particles, density, temperature,dimensionality (by chang<strong>in</strong>g the shape of the external trap), and even the strengthof <strong>in</strong>teractions between atoms us<strong>in</strong>g a technique called Feshbach resonance [6]. Bya simple modification of the external magnetic field, the effective <strong>in</strong>teratomic <strong>in</strong>teractioncan be tuned <strong>in</strong> the range of several orders of magnitude and this featuremakes cold atomic systems really unique. In more recent experiments, atoms aretrapped <strong>in</strong> periodic potentials, the so-called optical lattices [7]. In this way, it isnow possible to study, <strong>in</strong> a very clean setup, systems which are highly relevant <strong>in</strong>condensed matter physics, and which are not yet completely understood (with a notableexample of high-temperature superconductivity). For this reason, it is widelyaccepted that ultracold quantum gases represent Feynman’s quantum simulators [8].Intensive experimental progress and realization of the new ultracold phase of matterare strong stimuluses for further, <strong>in</strong>terdiscipl<strong>in</strong>ary theoretical research.The ma<strong>in</strong> subject of this <strong>thesis</strong> is a thorough understand<strong>in</strong>g of two <strong>in</strong>terest<strong>in</strong>gphysical scenarios for the manipulation of cold bosonic atoms, which were also thefocus of recent experimental studies. First, we have explored the phase diagram ofa rotat<strong>in</strong>g ideal bosonic gas <strong>in</strong> an anharmonic trap, while the second ma<strong>in</strong> topicdeals with nonl<strong>in</strong>ear features of collective modes excited by harmonic modulation of<strong>in</strong>teraction strength.On the way to accomplish this, after <strong>in</strong>troductory Chapter 1, <strong>in</strong> Chapter 2 wehave first worked out details of an efficient numerical method capable of provid<strong>in</strong>g


highly accurate <strong>in</strong>formation on energy levels of quantum systems. The precise <strong>in</strong>formationon energy spectra is necessary for the characterization of the phase diagramof a Bose-E<strong>in</strong>ste<strong>in</strong> condensate. Method that we have used is based on the exactdiagonalization of the time-evolution operator [9]. In order to optimally apply it,we have carefully analyzed numerical errors which arise for two reasons: numericalerrors which stem from the spatial discretization, as well as the errors due tothe approximative calculation of matrix elements. One of our ma<strong>in</strong> results is highlysuperior behavior of the dicretization error of the discretized evolution operator comparedto the common discretization error of the discretized Hamiltonian. Based onthe analytical and numerical considerations, we have shown that the diagonalizationof a time-evolution operator exhibits a non-perturbatively small discretization error,which vanishes exponentially with 1/∆ 2 , where ∆ is the discretization spac<strong>in</strong>g,while standard discretization <strong>in</strong>troduces errors polynomial <strong>in</strong> ∆. This is the ma<strong>in</strong>reason that makes the diagonalization of the time-evolution operator the preferredmethod. The ma<strong>in</strong> difficulty <strong>in</strong> the application of the method - precise calculationof the matrix elements of time-evolution operator (transition amplitudes) can be directlyresolved us<strong>in</strong>g previously developed effective action approach [10, 11], whichyields transition amplitudes as high-order expansions <strong>in</strong> the short time of propagation.The efficiency of this method has been demonstrated on several one- andtwo-dimensional models.In Chapter 3 we have used the described numerical method to explore the phasediagram of rotat<strong>in</strong>g bosons <strong>in</strong> a non-standard external potential. Widely used conf<strong>in</strong>ementsare harmonic traps and many details of Bose-E<strong>in</strong>ste<strong>in</strong> condensation <strong>in</strong>such traps are already well understood. Rotation of a quantum gas is one way toreach strongly correlated phases [7] and is therefore highly relevant. One of the consequencesof rotation is the appearance of the deconf<strong>in</strong><strong>in</strong>g centrifugal component <strong>in</strong>the potential, which <strong>in</strong> the case of a very fast rotation frequency (exceed<strong>in</strong>g the trapp<strong>in</strong>gfrequency) leads to the deconf<strong>in</strong>ement. In order to avoid this, recent experiment[12] <strong>in</strong>troduced an additional quartic potential to enhance trapp<strong>in</strong>g. Depend<strong>in</strong>g onthe value of the rotation frequency, the potential changes its shape from a simpleconvex one to the Mexican-hat-shaped potential. Us<strong>in</strong>g exact diagonalization of thetime-evolution operator, we have studied how the modification of the external trap<strong>in</strong>fluences properties of a Bose-E<strong>in</strong>ste<strong>in</strong> condensate, such as condensation temperature,equilibrium density distribution of atoms and the expansion time of the cloudafter it is released from the trap.


The effects of weak <strong>in</strong>teractions on the phenomenon of Bose-E<strong>in</strong>ste<strong>in</strong> condensationare subject of Chapter 4. This chapter reviews the Hartree-Fock descriptionof a bosonic system. Hartree-Fock is one of the mean-field approximations andwith<strong>in</strong> this framework we have described ultracold bosons at zero temperature, aswell as <strong>in</strong> the vic<strong>in</strong>ity of the Bose-E<strong>in</strong>ste<strong>in</strong> phase transition. At zero temperaturewe have rederived famous Gross-Pitaevskii equation [13, 14] - nonl<strong>in</strong>ear partial differentialequation, which governs the dynamics of the macroscopic wave functionof the condensate. For f<strong>in</strong>ite temperatures, we have scrut<strong>in</strong>ized several widely usedimplementations of the mean-field approximation, with the emphasis on their benefitsand their physical drawbacks. Approximations of this type are very often used<strong>in</strong> the <strong>in</strong>terpretation of the experimental data, which makes them highly relevant.A very recent experimental progress <strong>in</strong> the observation of beyond-mean-field effectsdemands further improvements of these standard tools.One of the basic methods to characterize phases of matter, both experimentallyand theoretically, are properties of their excitation spectra, with collective modesbe<strong>in</strong>g of special <strong>in</strong>terest. In ultracold gases, collective modes are usually excited us<strong>in</strong>gmodulation of the parameters of the external trap. In the recent experiment [15],however collective modes were excited us<strong>in</strong>g an alternative method - harmonic modulationof <strong>in</strong>teraction. In its essence the experimental method relies on a Feshbachresonancetechnique, which enables modulation of <strong>in</strong>teraction via a modulation ofthe external magnetic field. As an outcome of this dynamical protocol, oscillationsof the condensate size were <strong>in</strong>duced and measured. Depend<strong>in</strong>g on the value of theexternal modulation frequency, either a l<strong>in</strong>ear response or resonant large-amplitudeoscillations were obta<strong>in</strong>ed. S<strong>in</strong>ce the ma<strong>in</strong> underly<strong>in</strong>g equation is nonl<strong>in</strong>ear, <strong>in</strong> thecase of large-amplitude oscillations strong nonl<strong>in</strong>ear effects are expected. In Chapter5 we have performed numerical simulations of the system dynamics and identifiedma<strong>in</strong> nonl<strong>in</strong>ear features of the obta<strong>in</strong>ed excitation spectra: beside the basic modes,higher harmonics appear together with their l<strong>in</strong>ear comb<strong>in</strong>ations. Most prom<strong>in</strong>entnonl<strong>in</strong>ear effects are nonl<strong>in</strong>earity-<strong>in</strong>duced shifts <strong>in</strong> the frequencies of excited modes.In order to describe these results <strong>in</strong> an analytic way, we have developed perturbativeapproach where the small parameter is given by the modulation amplitude.The developed perturbative approach is based on the Po<strong>in</strong>caré-L<strong>in</strong>dstedt method,and gives analytical estimates for the shifts of the eigenfrequencies <strong>in</strong> the vic<strong>in</strong>ity ofresonances.


Keywords: cold quantum gases, Bose-E<strong>in</strong>ste<strong>in</strong> condensation, effectiveaction, exact diagonalization, perturbation theory, nonl<strong>in</strong>ear dynamicsField of Science: PhysicsResearch Area: Condensed matter physicsUDC number: 538.9


ContentsNomenclaturexviii1 Introduction 11.1 Foreword . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.2 Few basic facts on Bose-E<strong>in</strong>ste<strong>in</strong> condensation . . . . . . . . . . . . . 31.2.1 Non<strong>in</strong>teract<strong>in</strong>g bosonic gas <strong>in</strong> the harmonic trap . . . . . . . . 31.2.2 Experimental realization . . . . . . . . . . . . . . . . . . . . . 91.2.3 Interact<strong>in</strong>g bosons at low temperatures . . . . . . . . . . . . . 131.3 This <strong>thesis</strong> . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162 Properties of quantum systems viadiagonalization of transition amplitudes 192.1 Space-discretized Schröd<strong>in</strong>ger equation . . . . . . . . . . . . . . . . . 222.2 Discretization effects . . . . . . . . . . . . . . . . . . . . . . . . . . . 262.3 Effective actions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 332.4 Numerical results for one-dimensional models . . . . . . . . . . . . . 362.5 Numerical results for two-dimensional models . . . . . . . . . . . . . 472.6 Conclusions and outlook . . . . . . . . . . . . . . . . . . . . . . . . . 553 Thermodynamics of a rotat<strong>in</strong>g ideal BEC 573.1 Numerical calculation of energy eigenvalues and eigenstates . . . . . . 623.2 F<strong>in</strong>ite number of energy eigenvalues and semiclassical corrections . . 663.3 Global properties of rotat<strong>in</strong>g BECs . . . . . . . . . . . . . . . . . . . 703.3.1 Condensation temperature . . . . . . . . . . . . . . . . . . . . 703.3.2 Ground-state occupancy . . . . . . . . . . . . . . . . . . . . . 743.3.3 Comparison with semiclassical approximation . . . . . . . . . 753.4 Local properties of rotat<strong>in</strong>g BECs . . . . . . . . . . . . . . . . . . . . 773.4.1 Density profiles . . . . . . . . . . . . . . . . . . . . . . . . . . 77xv


CONTENTS3.4.2 Time-of-flight graphs for BECs . . . . . . . . . . . . . . . . . 793.4.3 Overcritical rotation . . . . . . . . . . . . . . . . . . . . . . . 803.5 Conclusions and outlook . . . . . . . . . . . . . . . . . . . . . . . . . 824 Mean-field description of an <strong>in</strong>teract<strong>in</strong>g BEC 844.1 Gross-Pitaevskii equation . . . . . . . . . . . . . . . . . . . . . . . . . 884.2 F<strong>in</strong>ite-temperature properties of a BEC . . . . . . . . . . . . . . . . . 904.2.1 Almost-ideal model . . . . . . . . . . . . . . . . . . . . . . . . 914.2.2 Semi-ideal model . . . . . . . . . . . . . . . . . . . . . . . . . 924.2.3 TFSC model . . . . . . . . . . . . . . . . . . . . . . . . . . . 944.2.4 GPSC model . . . . . . . . . . . . . . . . . . . . . . . . . . . 954.2.5 Calculation of the condensation temperature . . . . . . . . . . 974.3 Experimental assessment of different models . . . . . . . . . . . . . . 1004.4 Conclusions and outlook . . . . . . . . . . . . . . . . . . . . . . . . . 1025 Nonl<strong>in</strong>ear BEC dynamics by harmonicmodulation of s-wave scatter<strong>in</strong>g length 1045.1 Variational description of low-ly<strong>in</strong>g modes . . . . . . . . . . . . . . . 1075.2 Harmonic modulation of the s-wave scatter<strong>in</strong>g length: experiment . . 1115.3 Harmonic modulation of the s-wave scatter<strong>in</strong>g length:theoretical framework . . . . . . . . . . . . . . . . . . . . . . . . . . . 1145.4 Spherically-symmetric BEC . . . . . . . . . . . . . . . . . . . . . . . 1175.4.1 Po<strong>in</strong>caré-L<strong>in</strong>dstedt method . . . . . . . . . . . . . . . . . . . . 1215.4.2 Results and discussion . . . . . . . . . . . . . . . . . . . . . . 1235.5 Axially-symmetric BEC . . . . . . . . . . . . . . . . . . . . . . . . . 1265.5.1 Po<strong>in</strong>caré-L<strong>in</strong>dstedt method . . . . . . . . . . . . . . . . . . . . 1265.5.2 Results and discussion . . . . . . . . . . . . . . . . . . . . . . 1295.6 Conclusions and outlook . . . . . . . . . . . . . . . . . . . . . . . . . 1326 Summary 134A Numerical solution of the GP equation 137B Time-dependent variational analysis 142List of papers by Ivana Vidanović 144xvi


CONTENTSReferences 146CURRICULUM VITAE - Ivana Vidanović 158xvii


NomenclatureRoman Symbolsagk BLMNn(⃗r)N 0s-wave scatter<strong>in</strong>g length<strong>in</strong>teraction strength given by 4π 2 a/MBoltzmann constantspace cutoffatomic masstotal number of particlestotal density of particlesnumber of particles <strong>in</strong> the condensaten 0 (⃗r) condensate densityN thnumber of thermal particlesn th (⃗r) density of thermal particlesTtT 0 ctemperaturetimecondensation temperature of ideal gasV (⃗r) external trap potentialGreek Symbolsβ<strong>in</strong>verse temperature, 1/k B Txviii


CONTENTS∆λ Tdiscretization spac<strong>in</strong>gthermal de Broglie wavelength at temperature Tµ chemical potentialω x,y,ztrapp<strong>in</strong>g frequencyψ(⃗r) condensate wavefunctionOther SymbolsPlanck constant divided by 2πB n (µ, T) Bose-E<strong>in</strong>ste<strong>in</strong> distribution⃗p⃗rmomentumreal-space coord<strong>in</strong>ateAcronymsBEC Bose-E<strong>in</strong>ste<strong>in</strong> condensateGPHFTFGross-PitaevskiiHartree-FockThomas-FermiTOF time-of-flightxix


Chapter 1Introduction1.1 ForewordThe essential <strong>in</strong>gredients of the quantum mechanical theory are dual particle-wavenature of the matter, and the notion of identical particles. The quantum <strong>in</strong>dist<strong>in</strong>guishabilityof particles has a profound impact on the statistical properties. Whileparticles with a half-<strong>in</strong>teger sp<strong>in</strong> (fermions) try to avoid each other due to the Pauliexclusionpr<strong>in</strong>ciple, particles with an <strong>in</strong>teger sp<strong>in</strong> (bosons) do not exhibit such restrictionsand are actually, as a consequence of the m<strong>in</strong>imal energy pr<strong>in</strong>ciple, try<strong>in</strong>gto occupy the same s<strong>in</strong>gle-particle ground state. These effects are captured by twodifferent probability distributions that describe the thermal equilibrium <strong>in</strong> the twotypes of physical systems. The Fermi-Dirac distribution for fermions was orig<strong>in</strong>allyderived <strong>in</strong> 1926 by its two authors <strong>in</strong>dependently, while try<strong>in</strong>g to <strong>in</strong>troduce quantizationconcepts <strong>in</strong>to an ideal gas of particles obey<strong>in</strong>g the Pauli exclusion pr<strong>in</strong>ciple.The Bose-E<strong>in</strong>ste<strong>in</strong> distribution for bosons was <strong>in</strong>troduced <strong>in</strong> 1924 by the jo<strong>in</strong>t effortof Bose and E<strong>in</strong>ste<strong>in</strong> [1, 2]. At the time, it was a miss<strong>in</strong>g piece of knowledge for thecomplete explanation of the Planck’s law of black body radiation.In the high-temperature limit, both distributions are well approximated by acommon Maxwell-Boltzmann distribution. The condition necessary for the effectsof the quantum statistics to become observable can be roughly estimated as follows.The non<strong>in</strong>teract<strong>in</strong>g particles of the mass M affect each other if their thermal deBroglie wavelength λ T at temperature T,√2πλ T =2k B TM , (1.1)is comparable to the <strong>in</strong>ter-particle distance given by n −1/3 , where n is a typical particledensity. Naturally, the notion of the “high temperature” and “low temperature”depends on the features of the considered system. For example, electrons <strong>in</strong> a typical1


metal at room temperature (300K) are quite accurately described by the Fermi-Diracdistribution, mak<strong>in</strong>g the Fermi-Dirac statistics widely present <strong>in</strong> our everyday experience.Orig<strong>in</strong>ally, the Bose-E<strong>in</strong>ste<strong>in</strong> statistics was developed for the photons.Immediately after it was derived it was suggested that <strong>in</strong> the low-temperature limita macroscopic occupation of the s<strong>in</strong>gle-particle ground state should emerge. A phaseexhibit<strong>in</strong>g the macroscopic quantum phenomenon is denoted as the Bose-E<strong>in</strong>ste<strong>in</strong>condensate (BEC). An important necessity for the system to develop this type ofbehavior is the conservation of a number of particles, and as the photons can bespontaneously created and annihilated, a BEC doesn’t appear <strong>in</strong> this type of a system.Although the concept of a BEC arises at the essential level of the theory, itmanaged to escape our clear observation until the first breakthrough experiments <strong>in</strong>1995 [3, 4]. Between 1924 and 1995 a BEC was ma<strong>in</strong>ly discussed <strong>in</strong> the relation tothe fasc<strong>in</strong>at<strong>in</strong>g low-temperature phenomena of superfluidity and superconductivitythat <strong>in</strong>clude macroscopic quantum effects. On the other hand, sometimes it waseven considered as an artifact of a non<strong>in</strong>teract<strong>in</strong>g description of a bosonic gas.To expla<strong>in</strong> why the concept of a BEC may seem elusive, we elaborate on str<strong>in</strong>gentrequirements which are prerequisites to experimentally accomplish Bose-E<strong>in</strong>ste<strong>in</strong>condensation. As already discussed, two relevant length scales are the de Brogliewavelength and the <strong>in</strong>ter-particle distance. For the first scale to be large enough,<strong>in</strong> experiments we need light particles at low temperatures. To make the secondlength scale comparable to λ T , the system of particles should be dense enough.However, the upper limit of the density is set by a requirement to rema<strong>in</strong> close tothe non<strong>in</strong>teract<strong>in</strong>g gas description. Strong <strong>in</strong>teractions could spoil the observationof a BEC phenomena, by modify<strong>in</strong>g the ground-state of the system significantly.It is well known that at low temperatures most of the elements are <strong>in</strong> the solid orliquid state, thus the non<strong>in</strong>teract<strong>in</strong>g gas description loses its relevance. Eventually,experiments [3, 16] reached the regime of a dilute weakly <strong>in</strong>teract<strong>in</strong>g gas of alkaliatoms, with typical densities of atomic clouds <strong>in</strong> the range 10 18 m −3 − 10 21 m −3 (sixorder of magnitudes lower than the density of air) and typical temperatures of theorder of nanokelv<strong>in</strong>. The hallmarks of the Bose-E<strong>in</strong>ste<strong>in</strong> condensation were unequivocallydiscerned <strong>in</strong> this range of parameters and the importance of the experimentalrealization was recognized by a Nobel prize for physics <strong>in</strong> 2001.With these experimental achievements, the w<strong>in</strong>dow <strong>in</strong>to unexplored landscapesof the nature was opened. S<strong>in</strong>ce then, the subject has evolved <strong>in</strong>to a more generalresearch field of ultracold atoms and we will mention only some of the topics cur-2


ently explored to illustrate the versatility of possibilities. Once that the propertiesof a weakly <strong>in</strong>teract<strong>in</strong>g quantum gas were understood and explored to some extent,more complex and <strong>in</strong>terest<strong>in</strong>g phases of matter came to the focus of experimentalresearch. Progress <strong>in</strong> the last decade has led to the realization of optical latticeswhich practically allow simulations of condensed matter and other systems with<strong>in</strong>ultracold atoms framework, <strong>in</strong> the sense of Feynman’s quantum simulator [8]. Phasediagrams of strongly <strong>in</strong>teract<strong>in</strong>g quantum gases with long-range dipolar <strong>in</strong>teractionare currently explored. Beside the equilibrium properties, non-equilibrium and dynamicalphenomena <strong>in</strong> these systems are also major research topic. Most recentadvances <strong>in</strong>clude the realization of non-Abelian gauge fields, as well as bosonic systemswith the sp<strong>in</strong>-orbit type of coupl<strong>in</strong>g. Beside bosonic atoms, experiments arenow performed with ultracold fermionic atoms and even with ultracold molecules(emergent field of ultracold chemistry). Even a photonic BEC has been producedrecently [17]. Of course, the excit<strong>in</strong>g experimental developments are led and closelyfollowed by theoretical studies. Some of those will be presented <strong>in</strong> this <strong>thesis</strong>.1.2 Few basic facts on Bose-E<strong>in</strong>ste<strong>in</strong> condensationTo set up the stage and to sharpen our <strong>in</strong>tuition on the facets of cold bosonicgases, first we review a textbook knowledge on the Bose-E<strong>in</strong>ste<strong>in</strong> condensation of anon<strong>in</strong>teract<strong>in</strong>g gas. Initial ideas about this type of phase transition were based onthe considerations of a homogenous gas of particles. However, as will be expla<strong>in</strong>edlater <strong>in</strong> this Chapter, all experimental realizations <strong>in</strong>clude an external conf<strong>in</strong><strong>in</strong>g trap.Most widely used is a harmonic trap and hence here we consider thermodynamicproperties of harmonically trapped non<strong>in</strong>teract<strong>in</strong>g bosons [18, 19].1.2.1 Non<strong>in</strong>teract<strong>in</strong>g bosonic gas <strong>in</strong> the harmonic trapThe average occupation of a s<strong>in</strong>gle-particle state of the energy E n at temperature T<strong>in</strong> the gas of non<strong>in</strong>teract<strong>in</strong>g particles is given by the Bose-E<strong>in</strong>ste<strong>in</strong> distribution asB n (µ, T) =1e β(En−µ) − 1 , (1.2)where β = 1/k B T is the <strong>in</strong>verse temperature and µ stands for the chemical potential.Formally speak<strong>in</strong>g, we use the grand canonical ensemble description and assume3


that the system can exchange both the energy and the particles with a reservoir. Byadjust<strong>in</strong>g the value of the chemical potential µ, the average total number of particlesN can be tuned to the desired value, correspond<strong>in</strong>g to the number of particles <strong>in</strong>the systemN = N(µ, T) =∞∑n=01e β(En−µ) − 1 . (1.3)From Eq. (1.3), by a careful <strong>in</strong>spection, we realize that µ <strong>in</strong>creases as the temperaturegets lower. Also, we notice that, <strong>in</strong> order to have a well def<strong>in</strong>ed probabilitydistribution, the condition µ ≤ E 0 has to be satisfied. It is this limitation that leadsto the conclusion that at the certa<strong>in</strong> temperature, for which µ ≈ E 0 , the number ofthermally occupied states,N th =∞∑n=11e β(En−E 0)− 1 , (1.4)becomes saturated. By lower<strong>in</strong>g the temperature further, the number of atoms thatcan be accommodated <strong>in</strong> the higher energy states decreases, yield<strong>in</strong>g a macroscopicoccupation N 0 of the ground state, N 0 ∼ N. For a f<strong>in</strong>ite number of particles,the condensation sets <strong>in</strong> once the saturation of a thermal states is reached and weformally def<strong>in</strong>e the condensation temperature T 0 c from the condition N = N th andµ = E 0 . This yieldsN =∞∑n=11e β0 c (En−E 0)− 1 , (1.5)where β 0 c = 1/k B T 0 c . Let us now focus on the harmonic trapp<strong>in</strong>g potentialV (⃗r) = 1 2 M(ω2 x x2 + ω 2 y y2 + ω 2 z z2 ) , (1.6)with well known energy levels E ⃗n = ( (n x + 1)ω 2 x + (n y + 1)ω 2 y + (n z + 1)ω 2 z), where⃗n = (n x , n y , n z ), and <strong>in</strong>dices n x , n y and n z take the <strong>in</strong>teger values 0, 1, 2, . . . The condensationtemperature can be found from Eq. (1.5), us<strong>in</strong>g µ = E ⃗ 0 = (ω 2 x+ω y +ω z ),and we get the follow<strong>in</strong>g implicit equation for Tc 0:N = ∑n x,n y,n z1e β0 c (ωxnx+ωyny+ωznz) − 1 . (1.7) 4


Summations <strong>in</strong> the last expression can be exactly performed only numerically, so <strong>in</strong>order to proceed further <strong>in</strong> an analytic way, we use a semiclassical approximation [19,5]. Practically, this means that we neglect the discretness of energy levels and replacesummations by the <strong>in</strong>tegrals over now cont<strong>in</strong>uous n x , n y and n z . Mathematicallyspeak<strong>in</strong>g, the approximation is justified if the values of the functions which wesum over do not vary significantly over the summation step. For the Eq. (1.7), thementioned condition translates <strong>in</strong>to the high-temperature limit k B T ≫ (E n+1 −E n ),where the thermal energy is larger than the typical spac<strong>in</strong>g of energy levels. Withthis simplification, we obta<strong>in</strong>:N ≈===∫ ∞ ∫ ∞ ∫ ∞0∞∑0∏m=1 j=x,y,z0∫ ∞0dn x dn y dn ze β0 c (ωxnx+ωyny+ωznz) − 1dn j e −β0 c mω jn j1 ∑ ∞1(βc 0)3 ω x ω y ω z m 3m=11(β 0 c) 3 ω x ω y ω zζ 3 , (1.8)where we <strong>in</strong>troduce the Bose function (the polylogarithm function) ζ α (x) = ∑ ∞ x nn=1 n αand the abbreviation ζ α ≡ ζ α (1) for the Riemann zeta function. From the lastexpression, we f<strong>in</strong>d that the condensation sets <strong>in</strong> for:k B Tc 0 = ¯ω N 1/3 , (1.9)ζ 1/33where ¯ω is a geometric mean ¯ω = (ω x ω y ω z ) 1/3 . Consequently, we dist<strong>in</strong>guish thephase with a macroscopic occupation of the ground state and denote it as the condensatephase, and the phase without a macroscopic value of N 0 that is designatedas the (normal) gas phase. However, the phase transition is well def<strong>in</strong>ed only <strong>in</strong> thethermodynamic limit [5, 20] and hence <strong>in</strong> the case of f<strong>in</strong>ite-size systems we refer toT 0 cas the condensation temperature <strong>in</strong>stead of the critical temperature.The BEC phase transition is usually depicted <strong>in</strong> the phase diagram show<strong>in</strong>g thecondensate fraction N 0 /N versus temperature. From Eq. (1.8), we <strong>in</strong>fer that <strong>in</strong> thecondensate phase, the number of atoms <strong>in</strong> the thermal component is proportional5


to T 3 and that the condensate fraction satisfies( ) 3N 0 TN = 1 − . (1.10)Tc0Next, we address the local properties of a BEC and analyze density profiles <strong>in</strong>two different phases. In the gas phase, the density profile is given by:n(⃗r) = n th (⃗r) =∞∑B n (µ, T)|ψ n (⃗r)| 2 , (1.11)n=0while <strong>in</strong> the condensate phase, us<strong>in</strong>g µ = E 0 , we haven(⃗r) = n 0 (⃗r) + n th (⃗r) = N 0 |ψ 0 (⃗r)| 2 +∞∑B n (E 0 , T)|ψ n (⃗r)| 2 , (1.12)where ψ n (⃗r) are the s<strong>in</strong>gle-particle eigenstates of the external trap potential. Summationsover the complete eigenspectrum of the harmonic trap potential cannot beperformed analytically, even though we know the energy levels explicitely. Veryoften, for the purpose of an analytical consideration, the density profile of the gasphase is calculated with<strong>in</strong> the semiclassical approximation where we use the classicalHamiltonian H(⃗r, ⃗p) = ⃗p 2 /2M + V (⃗r) and count the number of states <strong>in</strong> the phasespace volume d⃗rd⃗p as d⃗rd⃗p/(2π 3 ). With<strong>in</strong> this approximation, we calculaten th (⃗r) =∫n=1d⃗p 1(2π) 3 e β(H(⃗r,⃗p)−µ) − 1 = 1 ζλ 3 3/2 (e β(µ−V (⃗r)) ) . (1.13)TAn important prerequisite, which guarantees the validity of this approximation, isthat the de Broglie wavelength λ T of particles has to be smaller than the characteristiclength scale over which the external potential varies significantly [19]. Theapproximation cannot be used for the ground state, which has to be treated fullyquantum mechanically. In addition, we emphasize that, <strong>in</strong> order for the Eq. (1.13)to be well def<strong>in</strong>ed, the value of the chemical potential has to satisfy the conditionµ ≤ m<strong>in</strong> ⃗r V (⃗r) . (1.14)The requirement for the achievement of a BEC can be rephrased <strong>in</strong> terms ofcharacteristic length scales us<strong>in</strong>g Eq. (1.13). It is easy to show that, at the onset of6


Figure 1.1: The hallmark of the Bose-E<strong>in</strong>ste<strong>in</strong> condensation - a prom<strong>in</strong>ent densitypeak appears below the condensation temperature. Density profiles of the expandedcloud are shown at three different temperatures. From the left to the right we seebosonic cloud right above the condensation transition, just below the condensationtransition and <strong>in</strong> the regime with almost a pure condensate. The experimental resultis orig<strong>in</strong>ally presented <strong>in</strong> Ref. [3] and this figure is taken from Ref. [5].the BEC, the relationn(0)λ 3 T ≈ ζ(3/2) ≈ 2.61238 (1.15)holds. Not surpris<strong>in</strong>gly, this condition is very close to the <strong>in</strong>tuitive argument givenat the beg<strong>in</strong>n<strong>in</strong>g of the Chapter.The noticeable feature of the condensate phase <strong>in</strong> the harmonic trap, not present<strong>in</strong> the gas phase, is a prom<strong>in</strong>ent density peak located at the trap center that reflectsmacroscopically occupied ground state which has the symmetry of the trap potential,superimposed onto the broad thermal distribution. This is illustrated <strong>in</strong> Fig. 1.1.A bimodality of the density distribution is an important signature of the onset ofBose-E<strong>in</strong>ste<strong>in</strong> condensation.An important and convenient aspect of the harmonic trap is that the s<strong>in</strong>gleparticleground-state is localized both <strong>in</strong> the real and <strong>in</strong> the momentum space.Hence, beside static density profiles of the trapped atoms, another possibility forthe experimental differentiation of the phases is free expansion of the gas from thetrap. Initially, the gas is <strong>in</strong> the thermal equilibrium <strong>in</strong> the trap, and then suddenlythe trap is switched off and the gas is allowed to expend freely. To describe thethermal gas we use the semiclassical approximation. The density profile after the7


expansion time t is given by [21]= 1λ 3 Tn th (⃗r, t) = 1(2π) 3 ∫∏σ=x,y,z(d⃗pd⃗r 01e β(H(⃗r 0,⃗p)−µ)− 1 δ3 (⃗r − ⃗r 0 − ⃗p M t )) ( 1/21ζ1 + ωσt 2 2 3/2 e βµ− βM 2„«)ωx 2 x21+ωx 2 + ω2 y y2t2 1+ωy 2 + ω2 z z2t2 1+ωz 2t2 For long expansion times we f<strong>in</strong>d an approximate expression for the density distributionof a thermal gas <strong>in</strong> the form [21]n th (⃗r, t) ∼ 1 ( )βMr2βµ−ζλ 3 3/2 e 2t 2 . (1.16)TAs we see, although the <strong>in</strong>itial density distribution is anisotropic, after the expansionof the thermal cloud, we obta<strong>in</strong> an isotropic density profile. Another important th<strong>in</strong>gthat we learn from the result (1.16) is that expanded density profiles actually carry<strong>in</strong>formation on the <strong>in</strong>itial velocity distribution of the cloud. This becomes obviouswhen we rewrite Eq. (1.16) asn th (⃗r, t) ∼ 1 ( )ζλ 3 3/2 e − Mv22k B T. (1.17)TNow we compare this behavior with the expansion of a pure condensate (N = N 0 ) <strong>in</strong>the ground state of the trap (1.6). For the time evolution of a quantum mechanicalstate, we have∫ˆ⃗p2−itψ HO (⃗r, t) = 〈⃗r|e 2M |ψHO 〉 =⃗p2−itd⃗p 〈⃗r|⃗p〉e 2M 〈⃗p|ψHO 〉.=1π 3/4∏σ=x,y,z( ) −14(1 + itωσ ) −1 −2 e σ2 Mωσ 1−itωσ2 1+t 2 ωσ 2 . (1.18)Mω σFrom Eq. (1.18) we conclude that <strong>in</strong> the limit of long propagation times, the condensatewidths are given by (ω σ /M) 2t 1 [19]. Obviously, the expansion of the groundstate is spatially anisotropic and the aspect ratio of the cloud is <strong>in</strong>verted dur<strong>in</strong>gthe expansion. The behavior is substantially different from the isotropic expansionof the thermal cloud and can be used as a diagnostic tool for the presence of thecondensate <strong>in</strong> the system.Now that we have <strong>in</strong>troduced the theoretical notations, <strong>in</strong> the next subsection8


we discuss <strong>in</strong> some detail the experimental realization of a BEC.1.2.2 Experimental realizationAs already mentioned, it was a pursuit of the clear physical realization of a veryfundamental concept of a BEC that triggered an enormous amount of the experimentaleffort start<strong>in</strong>g <strong>in</strong> the eighties of the last century. Although today BECsof different atoms are readily produced <strong>in</strong> laboratories worldwide, the creation ofquantum degenerate atomic gases took many years dur<strong>in</strong>g which many experimentaltechniques at the forefront of technology were developed. In this subsection we givea basic description of today’s typical experimental setup with a very brief historicaloverview. The ma<strong>in</strong> reference we rely on is Ref. [21].Already <strong>in</strong> the early experimental phase of the quest for nanokelv<strong>in</strong> temperatures,possibility of stor<strong>in</strong>g atoms <strong>in</strong> any type of a vessel was ruled out. Instead, a properconfiguration of magnetic, optical or comb<strong>in</strong>ed magneto-optical potential is used asan external conf<strong>in</strong><strong>in</strong>g potential, usually called a trap. Atoms are trapped around thepotential m<strong>in</strong>imum due to <strong>in</strong>teraction of their <strong>in</strong>duced magnetic or electric dipolemoments with external field. In most cases, harmonic potentialV (⃗r) = 1 2 M(ω2 xx 2 + ω 2 yy 2 + ω 2 zz 2 ) , (1.19)is a reasonable approximation of the external trap potential. Different trap configurationshave been experimentally realized: for <strong>in</strong>stance, highly elongated traps(which can be considered as effectively one dimensional), or pancake shaped traps(effectively two-dimensional regime). Beside this common type of a harmonic conf<strong>in</strong>ement,periodic external potentials (optical lattices) are widely used <strong>in</strong> nowadaysexperiments. We will ma<strong>in</strong>ly consider an axially-symmetric harmonic trap potential,with the radial trap frequency ω x = ω y = ω. The usual form of the potential isV (⃗r) = 1 2 Mω2 (ρ 2 + λ 2 z z2 ) , (1.20)where we <strong>in</strong>troduce the trap aspect ratio λ z = ω z /ω. Typical values of the trapfrequencies are of the order of 2π × (10 − 100)Hz, with a typical length scale of theorder of several microns.The <strong>in</strong>itial attempts to realize a BEC, focused on the sp<strong>in</strong>-polarized hydrogen.The hydrogen was s<strong>in</strong>gled out among other atom types s<strong>in</strong>ce it rema<strong>in</strong>s <strong>in</strong> the9


gas phase even at very low temperatures. Correspond<strong>in</strong>gly, many techniques wereorig<strong>in</strong>ally developed for the hydrogen samples. Yet, it turned out that alkali atoms,such as Rb, Li, Na, K, have several advantages over hydrogen. First, they can becooled us<strong>in</strong>g laser cool<strong>in</strong>g techniques [21, 22] by the commercially available laserswhose wavelengths correspond to the transitions between energy levels of alkaliatoms. Second, stronger elastic collisions allow an improved cool<strong>in</strong>g rate <strong>in</strong> theprocess of evaporative cool<strong>in</strong>g [21, 23]. Evaporative cool<strong>in</strong>g can be performed <strong>in</strong>several ways. Usually it is implemented by lower<strong>in</strong>g the trap depth. In this way,atoms with more than average energy are removed from the trap, hence allow<strong>in</strong>g therema<strong>in</strong><strong>in</strong>g atoms to equilibrate at lower temperature. Another way is to performthe radio frequency forced evaporation by transferr<strong>in</strong>g <strong>in</strong>ternal state of energeticatoms <strong>in</strong>to the untrapped configuration. By vary<strong>in</strong>g the used frequency, the f<strong>in</strong>alvalue of the temperature can be controlled. One of the crucial steps that f<strong>in</strong>ally ledto the achievement of the high phase-space density necessary for the observation ofa BEC was precisely the comb<strong>in</strong>ation of the laser cool<strong>in</strong>g and evaporative cool<strong>in</strong>gtechniques.Another vital experimental aspect is the characterization of a BEC once it hasbeen produced. In typical experiments, the sample has the size from 10 3 to 10 6bosonic alkali atoms, trapped <strong>in</strong> the external field <strong>in</strong> the space volume of (100µm) 3 ,with a density <strong>in</strong> the range 10 18 −10 21 m −3 , and the temperature of 100 nK. How thesystem can be probed to confirm that it really represents a BEC and has propertiespredicted theoretically? As usual, the probe has to be “gentle” enough not to perturbthe system significantly, hence cold bosonic gases are ma<strong>in</strong>ly probed optically. Thereare several widely used techniques, and we mention the most important ones.From early experiments until today, the absorption imag<strong>in</strong>g rema<strong>in</strong>s the mostused tool for characterization of a BEC. The sample is irradiated by a laser beamresonant with an <strong>in</strong>ternal atomic transition and the shadow of the sample is measuredby a CCD camera. Such a procedure may not produce the desired resultsalways, s<strong>in</strong>ce the image can be blacked out by the high density of the sample. Forthis reason, once the BEC regime has been reached, the trap is usually turned off,the cloud is allowed to expand freely for several seconds, and then the absorptionimag<strong>in</strong>g is performed. The procedure is denoted as the time-of-flight (TOF) measurement.In general, the quantity which is measured <strong>in</strong> this way is the opticaldensity of the sample (responsible for the absorption). It is proportional to thecolumn particle density, which can be determ<strong>in</strong>ed by this measurement technique.10


In the first papers [3, 4], the TOF approach was used to prove the presence of thecondensate <strong>in</strong> the sample based on the signatures discussed <strong>in</strong> subsection 1.2.1 forthe non<strong>in</strong>teract<strong>in</strong>g gas <strong>in</strong> the harmonic trap. Time-of-flight images revealed theemergence of density peaks on top of broad thermal distributions. Also, the peak <strong>in</strong>the density became more and more prom<strong>in</strong>ent as the temperature decreased further.In addition, the measured velocity distribution demonstrated stark contrast <strong>in</strong> theregimes above and below condensation temperature, be<strong>in</strong>g isotropic <strong>in</strong> the formercase and highly anisotropic <strong>in</strong> the latter case, reflect<strong>in</strong>g the anisotropy of the trap.From the measured density profiles, other quantities can be also estimated. Usually,non<strong>in</strong>teract<strong>in</strong>g gas model is used for the extraction of the sample properties. Forexample, the temperature of the sample is determ<strong>in</strong>ed by fitt<strong>in</strong>g a function (1.16)to the experimental data for the thermal cloud.Obviously, the TOF measurements are destructive and make the observation ofthe condensate dynamics quite complicated. With this technique a real-time monitor<strong>in</strong>gof a dynamical process requires many steps: <strong>in</strong>itially, a BEC is produced,then <strong>in</strong> the certa<strong>in</strong> <strong>in</strong>terval time some dynamical process happens, and f<strong>in</strong>ally theBEC is destructively imaged. To obta<strong>in</strong> another piece of <strong>in</strong>formation on the dynamicalevolution, for another value of the time of evolution, the whole procedurehas to be repeated, start<strong>in</strong>g from the production of a BEC. Another difficulty arises<strong>in</strong> the <strong>in</strong>terpretation of experimental data. The non<strong>in</strong>teract<strong>in</strong>g gas model for theharmonic trap is simple enough to enable quantitative explanation, however, theaccurate description of the expand<strong>in</strong>g <strong>in</strong>teract<strong>in</strong>g gas at f<strong>in</strong>ite temperature is farmore complicated. Very early [16], dispersive imag<strong>in</strong>g techniques were utilized forgather<strong>in</strong>g <strong>in</strong>-situ <strong>in</strong>formation on the density profiles of cold gases <strong>in</strong> a nondestructiveway. One of the techniques used successfully for an <strong>in</strong>-situ imag<strong>in</strong>g of dense samplesis a phase-contrast imag<strong>in</strong>g. This procedure can be implemented through two differentexperimental protocols. In the first one, the light diffracted from the sampleis recorded, while <strong>in</strong> the second one, the <strong>in</strong>terference patterns of the phase-shifted<strong>in</strong>cident light and the diffracted signal are measured. In this way, <strong>in</strong>formation on thecomplex phase acquired by the non-resonant light <strong>in</strong> the atomic sample is extracted[24]. Aga<strong>in</strong>, the quantity which is measured is the optical density along the l<strong>in</strong>e ofsight, yet the measurement can be performed even for a very dense samples. Thistype of measurement allows multiple imag<strong>in</strong>g of the same cloud, hence it is muchmore convenient for study<strong>in</strong>g condensate dynamics.Up to now we have only considered a non<strong>in</strong>teract<strong>in</strong>g gas of particles, because11


this is the system that was orig<strong>in</strong>ally used to derive the Bose-E<strong>in</strong>ste<strong>in</strong> distribution.Of course, realistic systems require many-body approach <strong>in</strong>clud<strong>in</strong>g <strong>in</strong>teractions andthe non<strong>in</strong>teract<strong>in</strong>g description is only an approximation. Now we discuss the typeand strength of <strong>in</strong>teractions <strong>in</strong> the dilute atomic systems, and later <strong>in</strong> the <strong>thesis</strong> wewill expla<strong>in</strong> how they modify the non<strong>in</strong>teract<strong>in</strong>g picture.A predom<strong>in</strong>ant <strong>in</strong>teraction <strong>in</strong> dilute cold atom systems is a two-body <strong>in</strong>teraction,typically <strong>in</strong> the form of scatter<strong>in</strong>g of atoms. In this <strong>thesis</strong> we consider only a shortrange van der Waals <strong>in</strong>teraction between charge-neutral atoms. A dipole-dipole <strong>in</strong>teractionis also present, but usually can be neglected <strong>in</strong> the case of alkali atoms.However, it plays prom<strong>in</strong>ent role <strong>in</strong> recent experiments with 52 Cr, where the effectsof long-range dipolar <strong>in</strong>teraction become measurable [25]. Due to diluteness of thesystem, a typical atom-atom distance is larger then the effective <strong>in</strong>teraction range,and details of the short-range <strong>in</strong>teraction potential are not so important. In thelow-energy low-momentum limit, the potential can be parameterized by a s<strong>in</strong>gle parametera, represent<strong>in</strong>g the atomic s-wave scatter<strong>in</strong>g length. In the pseudopotentialdescription, the van der Waals <strong>in</strong>teraction is replaced by a contact potentialV <strong>in</strong>t (⃗r − ⃗r ′ ) = gδ(⃗r − ⃗r ′ ) , (1.21)where the <strong>in</strong>teraction strength is given byg = 4π2 aM . (1.22)The <strong>in</strong>teraction is repulsive for g > 0 and attractive <strong>in</strong> the opposite case. It is knownthat a BEC with attractive <strong>in</strong>teractions is unstable toward collapse if the particledensity is high enough [5]. In this <strong>thesis</strong>, we conf<strong>in</strong>e the discussion to the repulsively<strong>in</strong>teract<strong>in</strong>g BECs.In general, scatter<strong>in</strong>g properties of atoms depend on their <strong>in</strong>ternal states. Thisfact allows the adjustment of the s-wave scatter<strong>in</strong>g length by the control of the externalparameters of the system. More specifically, several atomic species exhibit theso-called Feshbach resonance - dependence of the scatter<strong>in</strong>g length on the externalmagnetic field B, which is given by the expressiona(B) = a BG(1 + ∆ BB − B ∞), (1.23)12


where a BG is the off-resonant scatter<strong>in</strong>g length, B ∞ is the resonance position and∆ B is the resonance width. Feshbach resonance is a very useful tool that allowsf<strong>in</strong>e and versatile tun<strong>in</strong>g of the <strong>in</strong>teraction strength <strong>in</strong> the cold atom systems overa range of several orders of magnitude from highly repulsive to highly attractiveregime. A phenomenon is well known <strong>in</strong> atomic and nuclear physics as Feshbach-Fano resonance [26, 27]. Beside the orig<strong>in</strong>al reference by Ties<strong>in</strong>ga et al. [28] whichanalyzed the benefits of us<strong>in</strong>g the resonance <strong>in</strong> the experiments with ultracold atoms,the underly<strong>in</strong>g theory has been discussed <strong>in</strong> the review paper by Ch<strong>in</strong> et al. [6] and<strong>in</strong> several textbooks [19, 29]. Here we give only a brief explanation, based on a factthat two atoms can <strong>in</strong>teract via an energetically open or closed channel. These twochannels are coupled and <strong>in</strong>clude Zeeman terms. By tun<strong>in</strong>g the external magneticfield it is possible to make a bound state of a closed channel resonant with the<strong>in</strong>come energy <strong>in</strong> the open channel. In that case the scatter<strong>in</strong>g length diverges, asgiven <strong>in</strong> Eq. (1.23) for B = B ∞ . The properties of Feshbach resonances are exploited<strong>in</strong> many experiments and one of the topics <strong>in</strong> this <strong>thesis</strong> will explore some relatedrecent experimental results.To summarize the subsection, the achievement of the BEC regime took a lot ofeffort and many techniques had to be developed for this purpose. However, once thishas been achieved, the cold atomic systems have become the cleanest experimentalsett<strong>in</strong>g for study<strong>in</strong>g macroscopic quantum phenomena. All parameters of thesesystems are highly controllable and tunable: the geometry, temperature, densityand even the type and the strength of <strong>in</strong>teraction. For this reason the field of ultracoldatoms is still <strong>in</strong> the process of strong expansion and cross-collaboration withother fields, and further new important <strong>in</strong>sights are expected.1.2.3 Interact<strong>in</strong>g bosons at low temperaturesAfter discuss<strong>in</strong>g physical characteristics of cold bosonic atoms, we are ready to writedown the Hamiltonian of the system:∫Ĥ =d⃗r(− ˆψ † (⃗r) 22M ∇2 ˆψ(⃗r) + V (⃗r) ˆψ† (⃗r) ˆψ(⃗r) + g )2 ˆψ † (⃗r) ˆψ † (⃗r) ˆψ(⃗r) ˆψ(⃗r) . (1.24)Here ˆψ † (⃗r) and ˆψ(⃗r) are bosonic field operators <strong>in</strong> the second quantized form, andsatisfy commutation relations [ ˆψ(⃗r), ˆψ(⃗r ′ )] = 0, [ ˆψ † (⃗r), ˆψ † (⃗r ′ )] = 0, [ ˆψ † (⃗r), ˆψ(⃗r ′ )] =δ(⃗r−⃗r ′ ). On the right hand side of Eq. (1.24) we have the k<strong>in</strong>etic energy term, poten-13


tial energy of particles <strong>in</strong> the external trap given by V (⃗r) and <strong>in</strong>teraction term whichstems from ∫ d⃗rd⃗r ′ V <strong>in</strong>t (⃗r − ⃗r ′ ) ˆψ † (⃗r) ˆψ(⃗r) ˆψ † (⃗r ′ ) ˆψ(⃗r ′ ) = g ∫ d⃗r ˆψ † (⃗r) ˆψ † (⃗r) ˆψ(⃗r) ˆψ(⃗r).Hamiltonian given by Eq. (1.24) without the external trap potential was studied<strong>in</strong> the early work of Bogoliubov <strong>in</strong> relation to superfluidity observed <strong>in</strong> 4 He [30].It represents the start<strong>in</strong>g po<strong>in</strong>t <strong>in</strong> theoretical studies of a BEC. Here we give anelementary exposition of ma<strong>in</strong> ideas of the Bogoliubov and, <strong>in</strong> parallel, we <strong>in</strong>troduceconcepts important for the description of an <strong>in</strong>teract<strong>in</strong>g BEC.In the homogenous case, a good quantum number of the system is the wavevector⃗ k. Therefore, we use the decompositionˆψ(⃗r) = 1v 1/2 ∑⃗ ke i⃗ k⃗râ⃗k , (1.25)where â ⃗k are annihilation operators <strong>in</strong> the occupation number basis, with s<strong>in</strong>gleparticleeigenfunctions <strong>in</strong> the form of plane waves. In the non<strong>in</strong>teract<strong>in</strong>g limit, weexpect macroscopic occupation N 0 of the ground state that we designate as |N 0 〉. Ifwe take <strong>in</strong>to account the exact relationsâ 0 |N 0 >= √ N 0 |N 0 − 1 >, â † 0|N 0 >= √ N 0 + 1|N 0 + 1 > , (1.26)and the fact that N 0 + 1 ≈ N 0 − 1 ≈ N 0 , we arrive at the follow<strong>in</strong>g approximationswhich replace operators â 0 and a † 0 with c-numbers: â 0 ≈ √ N 0 and â † 0 ≈ √ N 0 . Theerror made by us<strong>in</strong>g this approximation is of the order of [a 0 , a † 0] = 1, and can beneglected compared to the macroscopic value of N 0 . Said another way, we haveapproximated the expression (1.25) byˆψ(⃗r) ≈ ψ + δ ˆψ(⃗r) , (1.27)where ψ = (N 0 /v) 1/2 = n 1/20 is a classical field, and the operator δ ˆψ(⃗r) correspondsto quantum fluctuations around the classical value. The next step <strong>in</strong> the Bogoliubovapproach is to consider quantum fluctuations as be<strong>in</strong>g small and to keep only termsup to the second order <strong>in</strong> δ ˆψ. With these simplifications we can f<strong>in</strong>d an appropriatetransformation which converts the quadratic approximation of the Hamiltonian <strong>in</strong>tothe diagonal form. The f<strong>in</strong>al result is the Bogoliubov excitation spectrum of the14


system given by( ) ǫ Bog ⃗k = √ 2⃗ k 22M( 2⃗ k 22M + 2gn 0). (1.28)In the limit of large | ⃗ k|, the spectrum yields excitations of a non<strong>in</strong>teract<strong>in</strong>g gas. Collectivephenomena appear <strong>in</strong> the long wavelength limit where the collective phononmode is found [30]. The microscopic derivation of the excitation spectrum (1.28)fits well with the Landau’s phenomenological description of the superfluid. Anotherrelevant quantity that can be derived with<strong>in</strong> the Bogoliubov framework is thecondensate depletion due to <strong>in</strong>teractions at T = 0. The number of non-condensedparticles is proportional to √ na 3 , which for the case of strongly <strong>in</strong>teract<strong>in</strong>g systemsuch as liquid 4 He yields the depletion as high as 90%. For this reason, the superfluidityof 4 He is considered only as an <strong>in</strong>direct manifestation of the Bose-E<strong>in</strong>ste<strong>in</strong>condensation.In general, the relation between the condensation and superfluidity is a subtleone. It can be shown that the macroscopic occupation of the ground state (1.27)<strong>in</strong>troduces the off-diagonal long range order (ODLRO) <strong>in</strong>to the system [31]:lim 〈 ˆψ † (⃗r) ˆψ(⃗r ′ )〉 = n 0 . (1.29)|⃗r−⃗r ′ |→∞However, long range correlations that are the manifestation of a superfluidity may bepresent even without the condensation, as for <strong>in</strong>stance <strong>in</strong> two-dimensional systems[7].In order to put the phenomenon of Bose-E<strong>in</strong>ste<strong>in</strong> condensation <strong>in</strong>to a broadercontext, we emphasize that the decomposition (1.27) represents the spontaneousbreak<strong>in</strong>g of the U(1) symmetry related to the conservation of the number of particles<strong>in</strong> the system described by the Hamiltonian (1.24). Hence, ψ is the order parameterthat acquires nonzero value <strong>in</strong> the condensed phase [31, 32].Quantitative tests of the theoretical concepts <strong>in</strong>troduced <strong>in</strong> this subsection becamepossible only with the experimental realization of Bose-E<strong>in</strong>ste<strong>in</strong> condensation<strong>in</strong> dilute vapors of alkali atoms. However, the experiments <strong>in</strong>troduced an importantadditional feature - the <strong>in</strong>homogeneity of the system due to the trapp<strong>in</strong>g potential.The first theoretical study of a <strong>in</strong>teract<strong>in</strong>g bosonic system <strong>in</strong> a harmonic trap waspresented <strong>in</strong> Ref. [33] <strong>in</strong> relation to the early experiments that were try<strong>in</strong>g to producehydrogen BEC. S<strong>in</strong>ce then, different approaches have been used to describe the15


condensation of <strong>in</strong>teract<strong>in</strong>g bosonic gas <strong>in</strong> an external trap potential. In this <strong>thesis</strong>we will work <strong>in</strong> the mean-field framework, with an order parameter <strong>in</strong>troduced <strong>in</strong> asimilar manner as <strong>in</strong> Eq. (1.27). The exposition of the mean-field framework for thetrapped system is given <strong>in</strong> Chapter 4.Throughout the years, a lot has been learned about the phenomenon of a BEC bya powerful comb<strong>in</strong>ation of <strong>in</strong>genious experiments and equally <strong>in</strong>genious theoreticalmodel<strong>in</strong>g. Yet, despite the <strong>in</strong>tensive progress <strong>in</strong> the field, there are still many openquestions related to the fundamental topics, as will be <strong>in</strong>dicated throughout the<strong>thesis</strong>.1.3 This <strong>thesis</strong>The ma<strong>in</strong> objective of this <strong>thesis</strong> is a thorough analysis and understand<strong>in</strong>g of two<strong>in</strong>terest<strong>in</strong>g physical scenarios for the manipulation of cold bosonic atoms that haverecently came <strong>in</strong>to the focus of the experimental research. Namely, we will firstexplore the details of the phase diagram of a rotat<strong>in</strong>g ideal bosonic gas <strong>in</strong> an anharmonictrap. As a second topic, we will <strong>in</strong>vestigate the nonl<strong>in</strong>ear features of theexcitation of collective oscillation modes by a modulation of the <strong>in</strong>teraction strength.On the way to accomplish this, <strong>in</strong> Chapter 2 we work out the details of thenumerical method which is capable of provid<strong>in</strong>g us with a highly-accurate energyspectrum of a few-body system. As we already saw <strong>in</strong> the subsection 1.2.1, the accurate<strong>in</strong>formation on the energy spectrum of the system is required for the descriptionof a BEC phase transition. The method that we elaborate on is based on the exactdiagonalization of the short-time evolution operator and was <strong>in</strong>troduced earlier <strong>in</strong> asimplified form [9]. To understand the benefits of the method, we first analyze theerrors associated with space discretization of the time-evolution operator. Basedon analytical and numerical analysis, we show that the discretization error vanishesexponentially with 1/∆ 2 , where ∆ is the discretization spac<strong>in</strong>g. This nonperturbativebehavior highly outperforms polynomial errors <strong>in</strong> discretization spac<strong>in</strong>g ∆which arises <strong>in</strong> the common real-space discretization of the Hamiltonian. The keycomplexity of the method is the accurate calculation of the matrix elements whichare given by transition amplitudes. To address this requirement, we apply recently<strong>in</strong>troduced effective action approach [10, 11] for obta<strong>in</strong><strong>in</strong>g short-time expansion ofthe propagator to very high orders. We demonstrate high efficiency of the methodon several one- and two-dimensional models.16


Hav<strong>in</strong>g the efficient numerical method at our disposal, <strong>in</strong> Chapter 3 we studyproperties of a rotat<strong>in</strong>g ideal gas. The <strong>in</strong>troduction of the angular momentum <strong>in</strong>tothe system is one way of reach<strong>in</strong>g the highly correlated regime <strong>in</strong> the cold atomsetup. As will be expla<strong>in</strong>ed, beside other effects, a rotation effectively <strong>in</strong>troduces adeconf<strong>in</strong><strong>in</strong>g component <strong>in</strong>to the trap potential. Particularly, <strong>in</strong> the regime of a fastrotation, the gas may experience the complete deconf<strong>in</strong>ement. Hence an additionalquartic potential was used for the trapp<strong>in</strong>g <strong>in</strong> the experiments from Ref. [12], butthe <strong>in</strong>terest<strong>in</strong>g regime of fast rotation has not been completely understood. Us<strong>in</strong>gthe exact diagonalization of a time evolution operator, we study numerically Bose-E<strong>in</strong>ste<strong>in</strong> condensation <strong>in</strong> the modified external potential which is a comb<strong>in</strong>ation ofthe harmonic and quartic component. The shape of the potential changes fromconvex with a s<strong>in</strong>gle m<strong>in</strong>imum to the Mexican hat shape, depend<strong>in</strong>g on the rotationfrequency. We explore how the change of the trapp<strong>in</strong>g potential <strong>in</strong>fluences the phasediagram properties. We also calculate the density profiles of the gas and time-offlightpictures <strong>in</strong> different regimes and f<strong>in</strong>d that typical time-scales for free expansionare <strong>in</strong>creased by an order of magnitude <strong>in</strong> the delicate regime of fast rotation.In Chapter 4, we cont<strong>in</strong>ue and expand a brief exposition of subsection 1.2.3, anddiscuss several different mean-field frameworks for the description of properties of aweakly <strong>in</strong>teract<strong>in</strong>g BECs. First we present the zero temperature mean-field description.We neglect quantum fluctuations and assume that all atoms are condensed atT = 0. In that case, we show that BEC properties are captured by the effectivenonl<strong>in</strong>ear equation, the famous Gross-Pitaevskii equation [34, 19, 13, 14]. Then wemove to the study of f<strong>in</strong>ite-temperature mean-field models of BEC. The relevanceof this aspect is two-fold: on one hand, mean-field models are widely used for the<strong>in</strong>terpretation of experimental data, and on the other hand, from the conceptualpo<strong>in</strong>t of view, it turns out that different models suffer from different unphysicaldrawbacks. We review and compare the exist<strong>in</strong>g models by calculat<strong>in</strong>g density profileswith<strong>in</strong> different approximations. To illustrate the <strong>in</strong>fluence of weak <strong>in</strong>teractionson the Bose-E<strong>in</strong>ste<strong>in</strong> condensation, we re-derive the mean-field <strong>in</strong>teraction-<strong>in</strong>ducedshift of the condensation temperature.Chapter 5 deals with the collective excitations of BEC <strong>in</strong> the nonl<strong>in</strong>ear regime.Characteristics of a BEC can be probed by monitor<strong>in</strong>g its dynamical response tothe external perturbation. Usually, the ground-state BEC is produced and thenit is perturbed by a modulation of the external trap potential. A specific featureof the recent experiment [15] is the harmonic modulation of the s-wave scatter<strong>in</strong>g17


length via a Feshbach resonance, yield<strong>in</strong>g a time-dependent <strong>in</strong>teraction strengthg = g(t). An external driv<strong>in</strong>g frequency Ω is used for the modulation, and depend<strong>in</strong>gon its closeness to some of the condensate eigenfrequencies, either a resonant ornon-resonant behavior can be observed. This is a new venue for study<strong>in</strong>g nonl<strong>in</strong>eardynamical regime, s<strong>in</strong>ce the equation govern<strong>in</strong>g the condensate dynamics on themean-field level is nonl<strong>in</strong>ear, and large amplitude oscillations are readily produced<strong>in</strong> the resonant regime. By comb<strong>in</strong><strong>in</strong>g different analytical and numerical methods weanalyze how nonl<strong>in</strong>ear effects <strong>in</strong>fluence the properties of the excited collective modes,with implications on the <strong>in</strong>terpretation of the actual experimental data. Prom<strong>in</strong>entnonl<strong>in</strong>ear features, such as: mode coupl<strong>in</strong>g, higher harmonics generation, and significantshifts <strong>in</strong> the frequencies of collective modes are found and quantitativelyexpla<strong>in</strong>ed us<strong>in</strong>g an analytic perturbative approach.At the end of each chapter, we present possible future directions for extend<strong>in</strong>g thestudy of research topics of this <strong>thesis</strong>. F<strong>in</strong>ally, we summarize all results <strong>in</strong> Chapter6. Additional material and derivations are organized <strong>in</strong> Appendices.18


Chapter 2Properties of quantum systems viadiagonalization of transition amplitudesDeep <strong>in</strong> the condensate phase, features of cold atoms are determ<strong>in</strong>ed by low-ly<strong>in</strong>genergy levels: ground state and few excited states correspond<strong>in</strong>g to the thermalcloud. On the other hand, thermodynamic properties and details of the BEC phasetransition are determ<strong>in</strong>ed by the full energy spectrum. As we have seen <strong>in</strong> Chapter1, the exact calculation of the condensation temperature, even for an ideal bosonicgas, requires a summation over the whole energy spectrum of the system. Dueto its simplicity, the semiclassical approximation is widely used for this purpose.In this Chapter we work out details of a numerical method based on the exactdiagonalization of a time-evolution operator that allows us access to a very largenumber of numerically exact energy levels of few-body systems. Afterwards, <strong>in</strong>Chapter 3, we use the method to f<strong>in</strong>d the condensation temperature of the fastrotat<strong>in</strong>gideal gas <strong>in</strong> an anharmonic trap.In the standard operator formulation of quantum mechanics, the description ofa physical system is based on construct<strong>in</strong>g the Hamilton operator Ĥ. Properties ofquantum systems are then obta<strong>in</strong>ed by solv<strong>in</strong>g the correspond<strong>in</strong>g time-<strong>in</strong>dependentSchröd<strong>in</strong>ger equation,Ĥ|ψ〉 = E|ψ〉 . (2.1)Exact solutions can be found only for a very limited set of simple models. A widevariety of analytical approximation techniques has been developed <strong>in</strong> the past fortreatment of such problems. In addition, the last two decades have seen a rapidgrowth <strong>in</strong> the application of different numerical methods for solv<strong>in</strong>g the Schröd<strong>in</strong>gerequation. Approaches based on real-space discretization start from some given f<strong>in</strong>itedifferenceprescription. Such methods have been extensively studied <strong>in</strong> the past, andthe ma<strong>in</strong> difficulties follow from the f<strong>in</strong>ite-difference representations of the k<strong>in</strong>eticoperator.A numerical approach based on diagonaliz<strong>in</strong>g of the evolution operator, <strong>in</strong>tro-19


2. Diagonalization of Transition Amplitudesduced <strong>in</strong> Ref. [9], does not suffer from problems with the representation of differentialoperators on real-space grids, and has substantial advantages <strong>in</strong> practical applicationsto few-body problems. Effectively, <strong>in</strong> this way the problem is transferred fromthat of represent<strong>in</strong>g the k<strong>in</strong>etic operator on a real-space grid to the calculat<strong>in</strong>g ofcorrespond<strong>in</strong>g transition amplitudes. Detailed analysis of the errors associated withthe implementation of this approach is the ma<strong>in</strong> objective of this Chapter. It providesfull understand<strong>in</strong>g of the method and allows its optimal use, as well as furthersignificant improvements with<strong>in</strong> a generalized calculation scheme.The advantages of the method discussed here [9, 35, 36, 37] follow from two keyproperties. First, the objects be<strong>in</strong>g diagonalized are transition amplitudes, which arewell def<strong>in</strong>ed irrespective of discretization scheme, i.e. the exponential of the Hamiltonianeffectively regularizes the k<strong>in</strong>etic operator, mak<strong>in</strong>g possible representations ofthe evolution operator that do not depend on the space grid. Second, the successfuldiagonalization of the evolution operator exp(−tĤ) for any time of propagation timmediately gives the solution of the eigenproblem for the Hamiltonian. Thus, thetime of propagation <strong>in</strong> this approach is just an auxiliary parameter. Said anotherway, we use the time-dependent evolution operator to extract time-<strong>in</strong>dependent<strong>in</strong>formation regard<strong>in</strong>g the quantum system. If one could calculate transition amplitudesexactly, then the obta<strong>in</strong>ed results for the energy eigenproblem would notdepend on the time of propagation. However, <strong>in</strong> practical applications one uses someapproximation scheme to calculate the amplitudes, and <strong>in</strong> this case the precision ofthe obta<strong>in</strong>ed results for energy eigenvalues and eigenstates does depend on time t.The general applicability of the outl<strong>in</strong>ed method follows from the fact that one canuse short-time propagation amplitudes to obta<strong>in</strong> highly accurate results.In order to complete this numerical method and make it generally applicable, itis necessary to address the follow<strong>in</strong>g key questions:1. How to analytically estimate the effects of spatial discretization?2. How to optimize the choice of evolution time t, so as to m<strong>in</strong>imize errors?3. How to accurately calculate transition amplitudes?The authors <strong>in</strong> Ref. [9] have only briefly commented on the first two questions,and numerically determ<strong>in</strong>ed the values of parameters that can be used for precisecalculations of energy eigenvalues and eigenstates for several models. In this Chapterwe address the above questions, which have not been fully answered before. First we20


2. Diagonalization of Transition Amplitudespresent the method and notation, and identify the sources of the errors present <strong>in</strong>real-space discretization approaches. Then we analyze <strong>in</strong> detail the above questions1 and 2, and discuss the effects of discretization on the numerically calculated valuesof the observables for a given physical system. We analytically derive estimates forerrors stemm<strong>in</strong>g from space discretization coarseness, f<strong>in</strong>ite size effects, and choice ofthe evolution time parameter t. All the analytically derived results are numericallyverified to hold on several <strong>in</strong>structive models.Errors associated with the time of evolution parameter t (question 3 above) mustbe carefully taken <strong>in</strong>to account and may substantially limit the precision of numericalcalculation <strong>in</strong> the diagonalization method. This problem was not addressed atall <strong>in</strong> Ref. [9], but has been addressed recently [38, 39, 40, 41, 42] us<strong>in</strong>g various approaches.We significantly improve the method by apply<strong>in</strong>g the recently <strong>in</strong>troducedeffective action approach [43, 44, 10, 45, 46, 11] to completely resolve the problemformulated <strong>in</strong> question 3. We stress that use of higher-order effective actionsrepresents an efficient and numerically <strong>in</strong>expensive way to calculate transition amplitudes,and leads to many orders of magnitude <strong>in</strong>crease <strong>in</strong> precision of calculatedproperties of the system. We will demonstrate on several lower-dimensional modelshow use of higher-order effective actions significantly reduces numerical errors andsystematically improves the obta<strong>in</strong>ed energy eigenvalues and eigenstates.This Chapter gives a complete analysis of the method based on the diagonalizationof transition amplitudes, provid<strong>in</strong>g us with necessary analytical knowledge toestimate errors of all types associated with this method and to numerically very accuratelycalculate large numbers of energy eigenvalues and eigenstates. This <strong>in</strong>vitesvarious applications of the method to the study of few-body quantum systems, someof which are discussed throughout the <strong>thesis</strong>.The expressions written throughout the second and third section are, for compactnessof notation, for one particle <strong>in</strong> one dimension. Extension to more particlesand dimensions is straightforward, just as with the above short-time transitionamplitude. Note that we are work<strong>in</strong>g <strong>in</strong> imag<strong>in</strong>ary time, which is well suited fornumerical calculations and does not affect <strong>in</strong> any way calculated energy levels norother time-<strong>in</strong>dependent properties of the system. We have also set to unity <strong>in</strong> thisChapter.21


2. Diagonalization of Transition Amplitudes2.1 Space-discretized Schröd<strong>in</strong>ger equationIn the coord<strong>in</strong>ate representation the time-<strong>in</strong>dependent Schröd<strong>in</strong>ger’s equation takesthe form∫dy 〈x|Ĥ|y〉 〈y|ψ〉 = E 〈x|ψ〉 . (2.2)The standard way to numerically implement exact diagonalization is to go fromcont<strong>in</strong>uous coord<strong>in</strong>ates x to ones liv<strong>in</strong>g on a discrete space grid x n = n∆, where∆ is a given spac<strong>in</strong>g and n ∈ Z. Integrations <strong>in</strong> the above equation are performedus<strong>in</strong>g the simple rectangular quadrature rule, or some higher-order f<strong>in</strong>ite-differenceformula. This completes the transition to the space-discretized counterpart of thecont<strong>in</strong>uous theory, however, to represent this on a computer we still have to restrictthe <strong>in</strong>tegers n to a f<strong>in</strong>ite range. This is equivalent to <strong>in</strong>troduc<strong>in</strong>g a space cutoff L, orputt<strong>in</strong>g the system <strong>in</strong> a <strong>in</strong>f<strong>in</strong>itely high potential box. For example, the rectangularquadrature rule leads to the follow<strong>in</strong>g space-discretized Schröd<strong>in</strong>ger equationN∑cut−1m=−N cutH nm 〈m∆|ψ〉 = E(∆, L) 〈n∆|ψ〉 , (2.3)where H nm = ∆ · 〈n∆|Ĥ|m∆〉, N cut = [L/∆], and square brackets represent the<strong>in</strong>teger part of the argument. As a result, we have obta<strong>in</strong>ed a 2N cut × 2N cut matrixthat represents the Hamiltonian of the system. The eigenvalues of this matrix dependon the two parameters <strong>in</strong>troduced <strong>in</strong> the above discretization process: cutoffL and discretization step ∆. Cont<strong>in</strong>uous physical quantities are recovered <strong>in</strong> thelimit L → ∞ and ∆ → 0. The outl<strong>in</strong>ed procedure is very useful <strong>in</strong> deal<strong>in</strong>g withspatially localized physical problems, such as electronic structure calculations <strong>in</strong>semiconductor and polymer physics [47].The two approximations <strong>in</strong>volved <strong>in</strong> the discretization procedure, characterizedby parameters ∆ and L, are common steps <strong>in</strong> solv<strong>in</strong>g eigenproblems of Hamiltonians<strong>in</strong> e.g. electronic structure calculations [47], and as such have been extensivelyanalyzed. The imposed constra<strong>in</strong>t on the values of spatial coord<strong>in</strong>ates to the f<strong>in</strong>ite<strong>in</strong>terval (−L, L) is a valid approach for captur<strong>in</strong>g <strong>in</strong>formation on localized eigenstates.In this approximation the system is effectively surrounded by an <strong>in</strong>f<strong>in</strong>itelyhigh wall, and as the cutoff L tends to <strong>in</strong>f<strong>in</strong>ity, we approach the exact energy levelsalways from above, which is a typical variational behavior. Therefore, we designateerrors associated with the cutoff L as variational. The effects of the discretization22


2. Diagonalization of Transition Amplitudesstep ∆ are more complex, and follow from the fact that the k<strong>in</strong>etic energy operatorcannot be exactly represented on f<strong>in</strong>ite real-space grids. For example, a typicalnaive discretization of the k<strong>in</strong>etic energy operator gives <strong>in</strong> our notation the follow<strong>in</strong>gHamiltonian matrix elements [48]⎧⎪⎨ 1/∆ 2 + V (n∆) if n = mH nm = −1/(2∆⎪⎩2 ) if |n − m| = 10 otherwise.(2.4)Note that <strong>in</strong> the absence of a potential term V <strong>in</strong> the Hamiltonian, the above def<strong>in</strong>itioncorresponds to a tight-b<strong>in</strong>d<strong>in</strong>g model [48]. This prescription leads to numericalresults for eigenvalues which <strong>in</strong> the ∆ → 0 limit converge to the exact cont<strong>in</strong>uumvalues as ∆ 2 . The errors associated with this approach have non-variational behavior,i.e. the obta<strong>in</strong>ed results are not always upper bounds of the exact energy levels.Several papers discuss this issue and analyze the behavior of errors <strong>in</strong> the direct diagonalizationapproach (for more details, see Refs. [49, 50] and references there<strong>in</strong>).The state-of-the-art <strong>in</strong> this approach is a set of systematically improved prescriptionsfor discretization of the k<strong>in</strong>etic energy operator, which speeds up convergenceto the cont<strong>in</strong>uum limit to higher powers of ∆ 2 . However, with<strong>in</strong> this approach convergenceis always polynomial <strong>in</strong> ∆. Some recent results [51] also exist on extensionsof this approach that provide effective variational behavior of the discretized k<strong>in</strong>eticenergy operator.As outl<strong>in</strong>ed <strong>in</strong> the Introduction, we focus on an alternative approach, based onsolv<strong>in</strong>g the eigenproblem of the correspond<strong>in</strong>g transition amplitudes as proposed <strong>in</strong>[9]. The central equation isor <strong>in</strong> the discretized formN∑cut−1e −tĤ|ψ〉 = e −tE |ψ〉 , (2.5)m=−N cutA nm (t) 〈m∆|ψ〉 = e −t E(∆,L,t) 〈n∆|ψ〉 , (2.6)where A nm (t) = ∆·A(n∆, m∆; t) = ∆·〈n∆|e −tĤ|m∆〉. In this approach the time ofevolution t plays the role of an auxiliary parameter. This parameter is not relatedto the discretization, and <strong>in</strong> a cont<strong>in</strong>uous theory it does not affect the obta<strong>in</strong>edeigenvalues and eigenstates. However, <strong>in</strong> a discretized theory the numerically calcu-23


2. Diagonalization of Transition AmplitudesA0.30.20.10-4-3-2-1-10-212-3x 34 -40123y4Figure 2.1: Harmonic oscillator transition amplitude as a function of coord<strong>in</strong>ates xand y for t = 1.lated eigenvalues and eigenstates will necessarily depend on this parameter as well,as emphasized by the right-hand size of Eq. (2.6). Therefore, the orig<strong>in</strong>al problemis now transformed <strong>in</strong>to the eigenproblem of the matrix A nm (t), whose <strong>in</strong>dices takeall <strong>in</strong>teger values <strong>in</strong> the range −N cut ≤ n, m < N cut , where N cut = [L/∆].Fig. 2.1 shows how a typical transition amplitude, <strong>in</strong> this case that of a harmonicoscillator, depends on coord<strong>in</strong>ates x and y. The transition amplitude of a harmonicoscillator with the Hamiltonian ĤHO = ˆp 2 /2 + ˆx 2 /2 can be calculated analyticallyand is given by explicit the expression:A HO (x, y; t) =√ [ 12π s<strong>in</strong>h t exp − 1 ((x 2 + y 2 ) cosht − 2xy )] . (2.7)2 s<strong>in</strong>h tNote that the consideration is general s<strong>in</strong>ce non-trivial mass and frequency of theharmonic oscillator can be easily taken <strong>in</strong>to account by a simple rescal<strong>in</strong>g of thecoord<strong>in</strong>ate and momentum. As can be seen from the figure, transition amplitudesare spatially well localized. This is particularly simple to understand for the shorttimes of propagation that we consider for a general case <strong>in</strong> the external potential.For a very short propagation times, transition amplitudes can be roughly calculatedas:A(x, y; t) ≈ √ 1 e −(x−y)2 −tV ( x+y2t 2 ) . (2.8)2πtFrom the last expression, we see that the k<strong>in</strong>etic term exponentially localizes thetransition amplitude matrix to the vic<strong>in</strong>ity of the ma<strong>in</strong> diagonal. Similarly, the24


2. Diagonalization of Transition Amplitudespotential br<strong>in</strong>gs about exponential localization along the ma<strong>in</strong> diagonal around itsm<strong>in</strong>imum. The localization of dom<strong>in</strong>ant values of the transition amplitude to asmall area <strong>in</strong> the x − y plane gives practical justification for <strong>in</strong>troduction of spacecutoff L <strong>in</strong> this approach.80706050t = 0.125t = 0.040t = 0.020t = 0.015E k403020100-100 5 10 15 20 25 30 35 40 45kFigure 2.2: Eigenspectrum of a free particle <strong>in</strong> a box. Eigenvalues E k are given asa function of level number k. The solid l<strong>in</strong>e gives the exact parabolic dispersionE k = π 2 (k + 1) 2 /8L 2 , while the dashed l<strong>in</strong>e presents results calculated <strong>in</strong> the tightb<strong>in</strong>d<strong>in</strong>gapproximation. The graph also shows numerical results obta<strong>in</strong>ed by thediagonalization of transition amplitudes for different values of time of evolution t.All the numerical calculations are for L = 6 and ∆ = 0.25, hence N cut = L/∆ = 24.In the cont<strong>in</strong>uum theory, the transition amplitude eigenproblem is mathematicallyequivalent to the Schröd<strong>in</strong>ger equation. It is important to stress, however,that the procedure of space discretization <strong>in</strong>troduces important differences betweeneigenproblems (2.3) and (2.6). In particular, as we will show <strong>in</strong> the next section, theprocedure based on the diagonalization of transition amplitudes leads to much faster(non-polynomial) convergence. An illustration of the relation of these two calculationschemes is shown <strong>in</strong> Fig. 2.2 which compares the exact parabolic dispersion ofa free particle <strong>in</strong> a box with numerical calculations based on diagonalizations of theHamiltonian and of the transition amplitudes. From the figure we see that the timeparameter t <strong>in</strong> the transition amplitude approach plays an important role. Increaseof the evolution time t gives better agreement with the exact dispersion relation.25


2. Diagonalization of Transition Amplitudes2.2 Discretization effectsThe free-particle transition amplitudeA free (x, y; t) = 1 √2πte −(x−y)2 2t (2.9)satisfies the relation∫dxA free (x, y; t) = 1 . (2.10)The consequence of this is conservation of probability. In the space-discretizedanalogue of this model x = n∆, y = m∆, and the transition amplitude is A freenm (t) =∆ A free (n∆, m∆; t). Us<strong>in</strong>g the Poisson summation formula√∑ πe −αn2 =αn∈Z∑n∈Ze − π2α n2 , (2.11)we f<strong>in</strong>d that the space discretized free particle amplitude satisfies∑n∈ZA freenm(t) = ∑ e −2π2 ∆ 2 n2t ≈ 1 + 2e −2π2 ∆ 2 t . (2.12)n∈ZConservation of probability is thus obta<strong>in</strong>ed only <strong>in</strong> the cont<strong>in</strong>uum limit ∆ → 0.Note that the effect of discretization is non-perturbative <strong>in</strong> discretization step ∆,i.e. it is smaller than any power of ∆. The effect of discretization is also universal<strong>in</strong> that it holds for all models, s<strong>in</strong>ce the free particle transition amplitude is thedom<strong>in</strong>ant term <strong>in</strong> the short time expansion of the transition amplitude of a generaltheory.To show this explicitly we use the short time expansion of the transition amplitudeof a general theory [44] to show that∫dxA(x, y; t) = 1 √2πt∫dxe − 1 2t x2 ∑ lt l f l (x, y) , (2.13)where f 0 = 1 and the other functions f l are determ<strong>in</strong>ed by the potential and itsderivatives. Writ<strong>in</strong>g the even part of f l (x, y) as g(x 2 , y) we f<strong>in</strong>d∫dxA(x, y; t) = √ 1 ∑t l g l (2t 2 ∂ t , y) √ t . (2.14)tl26


2. Diagonalization of Transition AmplitudesSimilarly, us<strong>in</strong>g the above Poisson summation formula, we f<strong>in</strong>d∑∫A nm (t) −n∈ZdxA(x, y; t) = 2 √t∑lt l g l (2t 2 ∂ t , y) √ te −2π2 ∆ 2 t . (2.15)Perform<strong>in</strong>g the <strong>in</strong>dicated differentiations the right hand side becomes exp(− 2π2∆ 2 t) ·∑l h l(y)t l . One could now calculate the functions h l from the short time expansionsof f l . The p-level effective action gives a short time expansion, which is truncatedat order t p . As a result, ∑ l h l(y)t l is a polynomial <strong>in</strong> time of order p. The dom<strong>in</strong>antshort time behavior is thus given by the universal exponential term. As a consequence,the transition of a general model to its space discretized form is given by∑∫A nm (t) −n∈ZdxA(x, y; t) ∼ e −2π2 ∆ 2 t . (2.16)This universal and non-perturbatively small deviation from the cont<strong>in</strong>uum <strong>in</strong>dicatesthat one should center numerical calculation schemes on transition amplitudesrather than the Hamiltonian. By diagonaliz<strong>in</strong>g the transition amplitude for anytime of propagation t we obta<strong>in</strong> the energy eigenvalues and eigenfunctions∫dy A(x, y; t)ψ k (y) = e −tE kψ k (x) . (2.17)To solve this numerically we first discretize space with discretization step ∆, andsecond we <strong>in</strong>troduce a spatial cut-off L such that |x| < L. Amplitudes are now2N cut × 2N cut matrices whose diagonalization leads to 2N cut eigenstates ψ k andeigenvalues e −tEk(∆,L,t) .As we have seen, discretization <strong>in</strong>troduces a non-perturbatively small error <strong>in</strong>transition amplitudes proportional to exp (−2π 2 t/∆ 2 ). We should therefore expectthe discretization error for energy eigenvalues to beE k (∆, L, t) − E k ∼ − 1 t e−2π2 ∆ 2 t . (2.18)We have numerically <strong>in</strong>vestigated this for a diverse set of models and have shown theabove relation to hold <strong>in</strong> all cases. It is also illustrative to verify this for analyticallytractable models. Us<strong>in</strong>g the known analytical expressions for transition amplitudeand energy eigenstates for a free particle <strong>in</strong> a box [52, 53], as well as the Poissonsummation formula <strong>in</strong> Eq. (2.11), we f<strong>in</strong>d that the energy eigenstates of the space27


2. Diagonalization of Transition Amplitudesdiscretized model satisfyE k (∆, L, t) − E k = − 2 t e−2π2 ∆ 2 t cosh π2 (k + 1)tL∆, (2.19)| E 0 (∆, L, t) - E 0 |10 -20 110 -4010 -6010 -8010 -100Ht = 0.25t = 0.5t = 110 -310 -410 -510 -60.2 0.4 0.6 0.8 10.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1∆10 20 0 0.2 0.4 0.6 0.8 1| E k (∆, L, t) - E k |10 -20 110 -4010 -6010 -8010 -10010 -120∆ E 0∆ E 14∆ E 24∆ E 34∆ E 44∆ E 54Figure 2.3: Top: Plot of |E 0 (∆, L, t)−E 0 | for a free particle <strong>in</strong> a box as a function of∆ for different values of time of evolution t and L = 6. For comparison, we also plotthe correspond<strong>in</strong>g deviations of numerical results (designated by H) obta<strong>in</strong>ed us<strong>in</strong>gdirect diagonalization of the space-discretized Hamiltonian, def<strong>in</strong>ed by Eq. (2.4).Bottom: This plot shows how the deviations |E k (∆, L, t) − E k | depend on t forseveral energy levels k. The parameters used are L = 6, ∆ = 0.2. In both plots thedashed l<strong>in</strong>es represent discretization error estimates given <strong>in</strong> Eq. (2.19).t28


2. Diagonalization of Transition Amplitudes12010080t = 0.015t = 0.020t = 0.040t = 0.125E k60402000 10 20 30 40 50 60 70 80 90kFigure 2.4: Harmonic oscillator dispersion relation. The solid l<strong>in</strong>e gives the exactl<strong>in</strong>ear dispersion E k = k +1/2. The po<strong>in</strong>ts correspond to numerically calculated energyeigenvalues E k as function of level k. We show the results of the diagonalizationof transition amplitudes for several values of t. In this plot L = 12, ∆ = 0.25.where E k = π2 (k+1) 2and k = 0, 1, 2, . . . As expected, the universal term gives the8L 2dom<strong>in</strong>ant ∆ dependence. One obta<strong>in</strong>s similar analytical results for the case of theharmonic oscillator.The non-perturbatively small effect of spatial discretization is the reason whythe new method highly outperforms direct diagonalization of the Hamiltonian andleads to much smaller errors for the same size of discretization step ∆. In additionto this the free parameter associated with the method, the time of evolution t, canbe used to further m<strong>in</strong>imize errors. As illustrated <strong>in</strong> Fig. 2.2, while keep<strong>in</strong>g ∆ fixed,we can adjust time t to obta<strong>in</strong> much smaller errors and practically reproduce theexact spectrum of the theory. This is also evident <strong>in</strong> Fig. 2.3, where we see thatby adjust<strong>in</strong>g t, errors can be reduced by orders of magnitude for fixed value ofdiscretization step ∆.We next consider a harmonic oscillator. Fig 2.4 shows how the presented methodmay be used to obta<strong>in</strong> energy eigenvalues to high levels. The numerical calculationsagree well with the well known l<strong>in</strong>ear dispersion of the harmonic oscillator.Figs. 2.5(top) and 2.5(bottom) display respectively the ∆ and t dependence of thedeviations |E k (∆, L, t) − E k |, show<strong>in</strong>g agreement with the analytically derived estimateof the discretization error given <strong>in</strong> Eq. (2.18). In order to achieve such a29


2. Diagonalization of Transition Amplitudes| E 0 (∆, L, t) - E 0 |110 -2010 -40Ht = 0.005t = 0.01t = 0.02t = 0.03t = 0.0410 -1 0 0.2 0.4 0.6 0.8 110 -310 -60 10 -50.05 0.1 0.15 0.2 0.25 0.310 -10 1∆| E k (∆, L, t) - E k |10 -2010 -3010 -4010 -5010 -6010 -70∆ E 0∆ E 14∆ E 24∆ E 34∆ E 44∆ E 540.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09tFigure 2.5: Top: Plot of |E 0 (∆, L, t) − E 0 | for a harmonic oscillator as a function ofdiscretization step ∆ for different values of time of evolution t with L = 12. For acomparison, we also plot the correspond<strong>in</strong>g results (designated by H) obta<strong>in</strong>ed us<strong>in</strong>gdirect diagonalization of the space-discretized harmonic oscillator Hamiltonian.Bottom: Plot of the deviations |E k (∆, L, t) − E k | given as a function of time t forseveral levels k. The parameters used are L = 12, ∆ = 0.1. In both plots the dashedl<strong>in</strong>es correspond to the discretization error estimate for E 0 given <strong>in</strong> Eq. (2.19).high accuracy of numerical results as presented on all graphs, we have used theMATHEMATICA software package [54].Fig. 2.5(bottom) shows how the deviations |E k (∆, L, t) − E k | depend on t forseveral levels k. The plot corresponds to the harmonic oscillator but is typical30


2. Diagonalization of Transition Amplitudesof a general theory. The saturation of errors for large t comes about when thediscretization error, given by the universal estimate <strong>in</strong> Eq. (2.18), becomes smallerthan the error due to space cutoff L. Analytical estimates for cutoff error are givenat the end of this section. At this po<strong>in</strong>t we just mention that the f<strong>in</strong>ite size effectscan already be seen <strong>in</strong> Fig. 2.4, where for high values of level number k numericalresults start to move away from the l<strong>in</strong>ear dispersion characteristic of a harmonicoscillator to the parabolic dispersion characteristic of a box potential.We end the section by look<strong>in</strong>g at f<strong>in</strong>ite size effects, i.e. errors related to <strong>in</strong>troductionof space cutoff L. For any theory with non-trivial potential, the cutoff Lis artificially <strong>in</strong>troduced and it affects the obta<strong>in</strong>ed energy eigenvalues, as we havealready discussed. To estimate the effects of the cutoff, we first note that they areclosely related to the spatial extent of the potential V , as well as the spatial extentof eigenfunctions of the system: errors <strong>in</strong> the correspond<strong>in</strong>g energy eigenvalues canbe considered small only if the eigenstates ψ k (x) are well localized <strong>in</strong> the <strong>in</strong>terval|x| < L.The effects of space cutoffs have been previously studied for cont<strong>in</strong>uous-spacetheories [55, 56]. The shift <strong>in</strong> energy level E k (L) −E k is found to be positive <strong>in</strong> thiscase, and approximately given by the formula(∫ L) −1dxE k (L) − E k = C k (a), (2.20)a |ψ k (x)| 2where a is an appropriately chosen value of coord<strong>in</strong>ate x such that it is larger thanand well away from the largest zero of ψ k (x) but smaller than and well away from thespace cutoff L. The constant C k (a) depends on the normalization of eigenfunctionand the choice of parameter a. For example, the ground state has no zeros, and wecan always choose the value a = 0. In that case, constant C 0 (0) is given by(∫ L) −1C 0 (0) = dx |ψ 0 (x)| 2 , (2.21)−Lwhere we assume that the eigenfunction ψ 0 (x) is normalized, ∫ ∞−∞ dx |ψ 0(x)| 2 = 1.In practical applications, when we use diagonalization of the discretized transitionamplitudes, the errors <strong>in</strong> energy level will necessarily also depend on theparameter t and other discretization parameters. Here we give a simple estimateof ground energy errors that follows from the spectral decomposition of diagonal31


2. Diagonalization of Transition Amplitudes10 10 110 -10 5 6 7 8 9 10 11 12E k (∆, L, t) - E k10 -2010 -3010 -4010 -5010 -6010 -70∆ E 0 , t = 0.1∆ E 0 , t = 10∆ E 14 , t = 0.1∆ E 14 , t = 10Figure 2.6: Deviations E k (∆, L, t) − E k for a harmonic oscillator as a function ofspace cutoff L for different values of time of evolution t. The discretization step is∆ = 0.1. Solid th<strong>in</strong> l<strong>in</strong>es give the dom<strong>in</strong>ant behavior of Eq. (2.23). The dashedthick l<strong>in</strong>es correspond to the error estimate <strong>in</strong> Eq. (2.20).Lamplitudes. For large t we have A(x, x; t) ≈ |ψ 0 (x)| 2 e −E0t . Integrat<strong>in</strong>g this we f<strong>in</strong>dan approximate result for the ground energy of a system with cutoff LE 0 (L, t) ≈ − 1 ∫ Lt ln dxA(x, x; t) , (2.22)In the L → ∞ limit we recover the exact ground energy, so that a simple estimateof f<strong>in</strong>ite size effects on E 0 is given byE 0 (L, t) − E 0 ≈ 1 ∫dx |ψ 0 (x)| 2 . (2.23)t−L|x|>LAlthough the above equation is just a rough estimate of the errors <strong>in</strong>troduced by aspace cutoff L, Fig. 2.6 shows that it is <strong>in</strong> good agreement with numerical resultsfor the harmonic oscillator. In order to clearly demonstrate L-dependence of errors<strong>in</strong> this graph, we have used small value of the discretization step ∆, such thatdiscretization errors can be neglected. The dashed l<strong>in</strong>es <strong>in</strong> the figure represent errorestimates given by Eq. (2.20).Us<strong>in</strong>g the data from Fig. 2.6 we can now fully expla<strong>in</strong> the saturation of errorsobserved <strong>in</strong> Fig. 2.5(bottom). The value of the cutoff L used to obta<strong>in</strong> this data32


2. Diagonalization of Transition Amplitudeswas L = 12. As can be seen from Fig. 2.6, this value of the cutoff parameter yieldsan error of the order 10 −65 for the ground-state energy for t ∼ 0.1, and of the order10 −40 for energy eigenlevel E 14 . These values exactly correspond to the saturatederrors <strong>in</strong> Fig. 2.5(bottom).Although <strong>in</strong> the general case the eigenstates that come <strong>in</strong>to Eqs. (2.20) and(2.23) are not known, we can still use them <strong>in</strong> conjunction with other approximationtechniques to estimate f<strong>in</strong>ite size effects. We also see that, due to the largerspatial extent of higher energy eigenstates, the cutoff-related errors are m<strong>in</strong>imal forthe ground energy. Note however that one is not really <strong>in</strong>terested <strong>in</strong> the precisecalculation of f<strong>in</strong>ite size errors, but only needs to estimate the m<strong>in</strong>imal size of thecutoff L for which f<strong>in</strong>ite size effects are negligible. For that purpose one can useeither of the above approximate formulas.2.3 Effective actionsSo far we have considered only <strong>in</strong>tegrable models, i.e. models for which we know theexact transition amplitudes. As a result we have thus far encountered and analyzedonly two sources of errors: those associated with discretization step ∆ and cutoffL. The vast majority of models are not <strong>in</strong>tegrable. The outl<strong>in</strong>ed method is stillapplicable if one uses some approximation for calculat<strong>in</strong>g transition amplitudes.As we can see, the precise calculation of transition amplitudes is essential forpractical applications of this method. In Ref. [9] all calculations are based on thenaive approximation for transition amplitudesA (1) (x, y; t) ≈1V (x)+V (y)−t√(2πt)d e−(x−y)2 2t 2 , (2.24)which is correct only to order O(t), and is for this reason designated by A (1) . If oneuses the naive approximation for transition amplitudes, then times of propagationmust be very short for errors to be small enough. Practically, even for short times ofpropagation, such errors are always much larger than the errors due to discretization,and therefore significantly limit the applicability of the method. In addition tothis, the results obta<strong>in</strong>ed <strong>in</strong> the previous section on exactly solvable models suggestthat longer times of propagation generally give smaller errors <strong>in</strong> the diagonalizationapproach. The trade-off between these effects and its implications on numericalresults have been documented <strong>in</strong> [9].33


2. Diagonalization of Transition AmplitudesTo address this, <strong>in</strong> pr<strong>in</strong>ciple one can use Monte Carlo simulations [57, 58] tocalculate amplitudes A to high precision. Although this can effectively resolve theproblem <strong>in</strong> many cases, it is often numerically very expensive. More importantly,resort<strong>in</strong>g to the use of Monte Carlo practically limits further analytical approaches.We will <strong>in</strong>stead use the recently <strong>in</strong>troduced effective action approach [44, 10, 45, 46,11] that gives closed-form analytic expressions A (p) (x, y; t) for transition amplitudeswhich converge much faster,A (p) (x, y; t) = A(x, y; t) + O(t p+1 /t d/2 ) , (2.25)where p is an <strong>in</strong>teger number correspond<strong>in</strong>g to the order of the effective actionused. For a general many-body theory effective actions up to p = 10 have beenderived, while for a specific models much higher values can be obta<strong>in</strong>ed, e.g. for theanharmonic oscillator and other polynomial <strong>in</strong>teractions, for which effective actionshave been calculated up to p = 144. So, if p is high enough, it is sufficient that thetime of evolution is less than the radius of convergence of the above series (t < τ c ∼ 1)and errors <strong>in</strong> calculated values of transition amplitudes will be negligible. This isillustrated <strong>in</strong> Fig. 2.7 for the case of a quartic anharmonic oscillator. The use ofhigh-order expansion <strong>in</strong> the time of propagation of amplitudes will allow us to usetimes of evolution up to τ c , which are much longer than the typical times one canuse with the naive (p = 1) amplitudes. At the same time, the expansion up to veryhigh orders substantially decreases the errors associated with t, and may practicallyelim<strong>in</strong>ate them.The analytic expressions for higher-order approximations for transition amplitudesare based on the notion of effective actions, which are <strong>in</strong>troduced by cast<strong>in</strong>gthe solution of the time dependent Schröd<strong>in</strong>ger equation for the transition amplitude<strong>in</strong> the formA(x, y; t) =1√(2πt)d e−(x−y)2 2t −tW( x+y ,x−y;t) 2 , (2.26)where W(x, δ; t) is the effective potential, with the follow<strong>in</strong>g boundary behavior:lim W(x, δ; t) = V (x) . (2.27)t→0As shown previously, the effective potential W(x, δ; t) is regular <strong>in</strong> the vic<strong>in</strong>ity oft = 0, enabl<strong>in</strong>g us to represent it <strong>in</strong> the form of a power series <strong>in</strong> short time ofpropagation t. The coefficients <strong>in</strong> this series are functions of the potential and34


2. Diagonalization of Transition AmplitudesA (p) (0, 0, t)21.51p = 1p = 10p = 20p = 30p = 400.500 0.5 1 1.5 2 2.5 3 3.5 4tFigure 2.7: Transition amplitude A (p) (0, 0; t) as a function of the time of propagationt, calculated analytically us<strong>in</strong>g different levels p of the effective action. The plot isfor the quartic anharmonic potential V (x) = k 22x 2 + k 424 x4 , with parameters k 2 = 1,k 4 = 10.its derivatives. The truncation of the series for the effective potential up to ordert p−1 , designated by W (p−1) (x, δ; t), gives the expansion of the transition amplitudeaccurate to t p ,A (p) (x, y; t) =1−tW√(2πt)d e−(x−y)2 2t(p−1) ( x+y ,x−y;t) 2 . (2.28)The analytic expressions for higher-order effective actions therefore yield analyticapproximations for amplitudes with the convergence behavior given by Eq. (2.25).We emphasize that although the structure of the effective action solution form (2.26)is motivated by the path <strong>in</strong>tegral formalism, the expression for amplitudes obta<strong>in</strong>ed<strong>in</strong> the above approach conta<strong>in</strong> no <strong>in</strong>tegrals and can be used straightforwardly aslong as the time of propagation is below the radius of convergence of the short-timeseries expansion.For the exactly solvable case of a harmonic oscillator one f<strong>in</strong>ds that the radiusof convergence is τ c = π. The radius of convergence is simply the distance <strong>in</strong> thecomplex time pla<strong>in</strong> from the orig<strong>in</strong> to the nearest s<strong>in</strong>gularity of the propagator.For the harmonic oscillator the s<strong>in</strong>gularities are located at ±ikπ, k ∈ N. Theconsequence of these s<strong>in</strong>gularities is that the power series for the effective potential35


2. Diagonalization of Transition AmplitudesW(x, δ; t) converges only for t < τ c . It is often difficult to analytically determ<strong>in</strong>ethe radius of convergence of the short time expansion of the transition amplitude.However, numerically this is a very simple problem, s<strong>in</strong>ce outside of the radius ofconvergence the calculated approximative amplitudes rapidly tend to <strong>in</strong>f<strong>in</strong>ity (forlevels p for which the effective potential is not bounded from below; see Ref. [59]) orto zero with the <strong>in</strong>crease of p. From Fig. 2.7 we easily estimate radius of convergenceto be τ c ≈ 1 for a quartic anharmonic potential V (x) = k 22x 2 + k 424 x4 , with parametersk 2 = 1, k 4 = 10. Such numerical determ<strong>in</strong>ation of the radius of convergence for agiven level p is always done before practical use of the effective potential. Note thatwe are not <strong>in</strong>terested <strong>in</strong> the precise value of τ c , just <strong>in</strong> its rough estimate which willallow us to safely use times of propagation below τ c .To conclude the section, let us stress that the effective action approach canbe used only for sufficiently smooth potentials, i.e. those that have derivatives ofthe required order, correspond<strong>in</strong>g to the level p of effective action, as discussed <strong>in</strong>Ref. [44]. For potentials that do not fulfill this condition (e.g. stepwise potentials),the effective action approach cannot be directly used. However, one can replacethe orig<strong>in</strong>al potential with some of its smooth deformations, perform numericalcalculations, and at the end take the limit of the deformation parameter <strong>in</strong> whichthe orig<strong>in</strong>al potential is recovered. The numerical results obta<strong>in</strong>ed <strong>in</strong> such a waymust be carefully cross-checked us<strong>in</strong>g other methods.2.4 Numerical results for one-dimensional modelsIn this section we apply the approach outl<strong>in</strong>ed above to several d = 1 modelsand demonstrate its substantial advantages for numerical studies of eigenstates ofvarious physical systems. We numerically analyze all sources of errors present <strong>in</strong>this approach due to discretization parameters L and ∆, as well as the time ofpropagation parameter t. We present the obta<strong>in</strong>ed numerical results for energyeigenvalues and eigenstates. We also assess the quality of the obta<strong>in</strong>ed energy spectrathrough comparison with the semiclassical approximation for the density of states,which should be accurate at least for the higher regions of the spectrum.The first model we study is the quartic anharmonic oscillator with potentialV (x) = k 22 x2 + k 424 x4 . (2.29)36


2. Diagonalization of Transition Amplitudes10 150.05 0.1 0.15 0.2 0.25 0.3| E 0 (∆, L, t) - E 0 |10 1010 5110 -510 -1010 -1510 -2010 -25Ht = 0.005t = 0.01t = 0.02t = 0.03t = 0.0410 -110 -210 -310 -40.1 0.2 0.3Figure 2.8: Plot of |E 0 (∆, L, t) − E 0 | for an anharmonic oscillator (2.29) given as afunction of ∆ for different values of time of evolution t. The parameters used are L =6, k 2 = 1, and anharmonicity k 4 = 48. Dashed l<strong>in</strong>es correspond to the discretizationerror <strong>in</strong> Eq. (2.19). For comparison, we also plot the correspond<strong>in</strong>g deviations ofnumerical results (designated by H) obta<strong>in</strong>ed us<strong>in</strong>g direct diagonalization of thecorrespond<strong>in</strong>g space-discretized Hamiltonian.∆For this potential the effective actions have been previously derived up to p = 144[60], and here we will use various levels p to illustrate the dependence of errors onthe level p used <strong>in</strong> calculations. We will study the <strong>in</strong>terest<strong>in</strong>g regime of the strongcoupl<strong>in</strong>g k 4 , s<strong>in</strong>ce there are various other techniques that can be successfully usedfor small coupl<strong>in</strong>gs.Fig. 2.8 displays |E 0 (∆, L, t) − E 0 | as a function of discretization step ∆ for thecase of an anharmonic oscillator with potential (2.29). The parameters used <strong>in</strong> theplot are L = 6, k 2 = 1, and anharmonicity k 4 = 48. The transition amplitude matrixelements were calculated us<strong>in</strong>g p = 18 effective actions [44]. The high precision valuefor the exact ground energy that we compare to was calculated <strong>in</strong> Ref. [61]. As wecan see, even though we are deal<strong>in</strong>g with a relatively strong anharmonicity, thenumerically calculated values stay right on the dashed black l<strong>in</strong>es correspond<strong>in</strong>g tothe universal discretization error just as <strong>in</strong> the case of the previously considered<strong>in</strong>tegrable models. This is <strong>in</strong> complete agreement with our analytical derivation ofthe discretization error.As can be seen from Fig. 2.8 the numerical results clearly demonstrate that the37


2. Diagonalization of Transition Amplitudes∆-dependence of errors with<strong>in</strong> our calculation scheme highly outperforms the polynomialdependence <strong>in</strong> ∆ 2 obta<strong>in</strong>ed by the direct diagonalization of the Hamiltonian.This is true even for short times of propagation t. Although <strong>in</strong>teraction terms <strong>in</strong>the potential affect the numerical values of errors, diagonalization of the transitionamplitudes still substantially outperforms diagonalization of the Hamiltonian, andis the preferred method. This success is a consequence of the non-perturbative behaviorof the spatial discretization error with<strong>in</strong> this calculation scheme. This leadsus to the key conclusion that discretization parameters can be always optimized sothat presented approach of solv<strong>in</strong>g eigenvalue problem of space-discretized transitionamplitudes highly outperforms direct diagonalization of the space-discretizedHamiltonian. The cont<strong>in</strong>uum limit ∆ → 0 is far more easily approached <strong>in</strong> thefirst case and the correspond<strong>in</strong>g discretization errors are substantially smaller forthe same discretization coarseness. From the numerical po<strong>in</strong>t of view, as the valueof parameter ∆ directly determ<strong>in</strong>es the size of the matrix to be diagonalized, thecomputational cost for the same precision is significantly reduced.Further analysis of various errors <strong>in</strong> the ground energy calculation for parametersk 2 = 1, k 4 = 48 of the anharmonic potential (2.29) is given <strong>in</strong> Fig. 2.9. Thedependence of the error related to the <strong>in</strong>troduction of the space cutoff L is illustrated<strong>in</strong> Fig. 2.9(top), while Fig. 2.9(bottom) gives the dependence of ground energy errorson the time of propagation parameter t for various values of the discretization step ∆.On both graphs we see the results obta<strong>in</strong>ed with effective actions of different levelsp. Fig. 2.9(bottom) clearly shows that the errors due to the time of propagationare proportional to t p , as expected when we use the effective action of the level p.The errors <strong>in</strong> eigenvalues are of the same order as errors <strong>in</strong> calculation of <strong>in</strong>dividualmatrix elements, and for this reason we see the typical t p behavior of ground andhigher energy eigenvalues. It is already now evident that the use of higher-ordereffective actions <strong>in</strong>creases the accuracy of numerically calculated energy eigenstatesfor many orders of magnitude.The L-dependence of the error is analytically known [55, 56]. The saturation oferrors on the top graph for a given level p corresponds to a maximal precision thatcan be achieved with that p, i.e. denotes the value of L for which errors <strong>in</strong>troduced byother sources become larger than the error due to the f<strong>in</strong>ite value of the space cutoff.This can be easily seen if we comb<strong>in</strong>e the data from both graphs. For example, thelevel p = 9 effective action has the saturated value of the error of the order of10 −14 . For t = 0.02 we f<strong>in</strong>d that the error due to the time of propagation is of the38


2. Diagonalization of Transition Amplitudessame order if one uses sufficiently f<strong>in</strong>e discretization (∆ = 0.05). Therefore, thesaturation of errors on the left-hand graph are caused by the errors due to the timeof propagation. However, if one uses discretization which is not sufficiently f<strong>in</strong>e, thesaturation of errors can be also caused by the discretization effects. Such effects can|| E 0(p) (∆, L, t) - E 0exact10 -2 110 -410 -610 -810 -1010 -1210 -1410 -1610 -1810 -20p = 1p = 3p = 5p = 7p = 9p = 11p = 131 1.5 2 2.5 3 3.5L10 5 0.001 0.01 0.1|| E 0(p) (∆, L, t) - E 0exact10 -5 110 -1010 -1510 -20∆ = 0.05∆ = 0.10∆ = 0.2010 -25Figure 2.9: Deviations from the ground energy |E (p)0 (∆, L, t) − Eexact 0 | as a functionof the space cutoff L (top) and as a function of the time t (bottom). The groundenergy is obta<strong>in</strong>ed us<strong>in</strong>g different levels p = 1, 3, 5, 7, 9, 11, 13 (top to bottom) ofthe effective action for the quartic anharmonic potential, with parameters k 2 = 1,k 4 = 48, ∆ = 0.05, t = 0.02 on top graph, and L = 4 on bottom graph. Theexact ground energy E0 exact = 0.95156847272950001114693 . .. is taken from Ref. [61].Dashed l<strong>in</strong>es on the graph (b) correspond to the discretization error (2.19).t39


2. Diagonalization of Transition AmplitudesTable 2.1: Low-ly<strong>in</strong>g energy levels of the anharmonic quartic potential, obta<strong>in</strong>edby diagonalization us<strong>in</strong>g level p = 13 effective action. The parameters are k 2 = 1,k 4 = 48, L = 5, ∆ = 0.05, t = 0.01. For higher energy eigenvalues, errors areestimated by comparison with the diagonalization results obta<strong>in</strong>ed from higherordereffective actions, f<strong>in</strong>er discretizations, larger space cutoffs, and lower values ofthe propagation time t.k E k |∆E k | δE k0 0.9515684727295000111468(8) 5 × 10 −23 6 × 10 −231 3.292867821434465922691(67) 4 × 10 −22 2 × 10 −222 6.30388056744652609989(522) 2 × 10 −21 4 × 10 −223 9.72732317270370501553(448) 5 × 10 −21 5 × 10 −224 13.4812758360385893838(1489) 2 × 10 −20 2 × 10 −215 17.5141323992530709259(6206) 3 × 10 −20 2 × 10 −216 21.7909563917965158973(8744) 6 × 10 −20 3 × 10 −217 26.286125156056810490(92289) 2 × 10 −19 7 × 10 −218 30.979882837938369575(08213) 2 × 10 −19 8 × 10 −219 35.856438766665971146(24181) 3 × 10 −19 9 × 10 −21be also analytically estimated to be proportional to −2 exp(−2π 2 t/∆ 2 ) cosh(π 2 (k +1)/L∆)/t, as we have shown <strong>in</strong> the previous section.Table 2.1 gives low-ly<strong>in</strong>g energy eigenvalues of the anharmonic oscillator for aparticular choice of the parameters of the potential and discretization parameters.In pr<strong>in</strong>ciple, one can achieve arbitrary precision by the use of appropriately chosendiscretization parameters. Of course, for arbitrary precision calculations one has touse one of the software packages able to support such calculations. For example,we have used MATHEMATICA [54] <strong>in</strong> order to be able to achieve high-precisionresults presented on the above graphs. The important conclusion is that even for verymoderate values of discretization parameters, the use of higher-order effective actionsleads to very small errors, which may be practically implemented with m<strong>in</strong>imalcomput<strong>in</strong>g resources.The analysis of errors such as the one presented <strong>in</strong> Fig. 2.9 is sufficient to estimateoptimal values of discretization parameters. In general, for a desired numericalprecision of energy eigenvalues, the optimal values of parameters are chosen so thatall types of errors are approximately the same. The overall error is always dom<strong>in</strong>atedby the largest of all errors, and therefore it is optimal to have all errors of the sameorder of magnitude.40


2. Diagonalization of Transition AmplitudesFor specific calculations one can have additional constra<strong>in</strong>ts. For example, if oneis <strong>in</strong>terested only <strong>in</strong> energy eigenvalues, then the optimal parameters are obta<strong>in</strong>edby m<strong>in</strong>imiz<strong>in</strong>g all errors and m<strong>in</strong>imiz<strong>in</strong>g the ratio N cut = L/∆, which correspondsto the size of the transition operator matrix S = 2N cut that needs to be numericallydiagonalized. The m<strong>in</strong>imization of N cut is performed <strong>in</strong> order to m<strong>in</strong>imizecomputation time needed for the diagonalization, which roughly scales as Ncut 3 .On the other hand, if one is <strong>in</strong>terested <strong>in</strong> details of energy eigenfunctions, then itmight be necessary to have a fixed small value for the discretization step ∆, whichwill allow all features of eigenstates to be visible. This is especially important forstudies of higher energy eigenfunctions which e.g. have many nodes, and <strong>in</strong> orderto study them it is necessary to have sufficient spatial resolution. In such case, thevalue of ∆ is fixed and other parameters are chosen so as to m<strong>in</strong>imize the errors toa desired value. For example, with the discretization step of the order ∆ = 10 −3 wehave been able to accurately calculate several hundreds energy eigenfunction of thequartic anharmonic oscillator.Table 2.2 gives eigenvalues of the double-well potential, obta<strong>in</strong>ed from the quarticanharmonic potential (2.29) by sett<strong>in</strong>g the constant k 2 to some negative value. Ascan be seen, numerically obta<strong>in</strong>ed energy eigenvalues have the precision similar to theprevious case of the quartic potential without symmetry break<strong>in</strong>g. The double wellbehavior of the potential does not present any obstacle <strong>in</strong> its numerical treatmentby this method.Another situation <strong>in</strong> which one might be <strong>in</strong>terested to keep the ratio N cut = L/∆,i.e. the size of the space-discretized evolution operator matrix as large as possibleis when a large number of energy eigenlevels is needed. The number of energyeigenvalues that can be calculated by the diagonalization is limited by the size ofthe matrix S = 2N cut . Usually the highest energy levels cannot be used due tothe accumulation of numerical errors, and therefore one needs to have a matrix ofsufficient size <strong>in</strong> order to study energy spectra. In such cases it is necessary to usehighly optimized libraries for numeric diagonalization. We have implemented theeffective actions as a C programm<strong>in</strong>g language code [60] and used LAPACK [62]library for numeric diagonalization to calculate large number of energy eigenvaluesand eigenfunctions.Even when one uses such a sophisticated tool, the highest eigenvalues cannotbe used due to accumulation of numerical errors. In order to assess the quality ofthe obta<strong>in</strong>ed results for higher energy eigenstates, it is necessary to compare the41


2. Diagonalization of Transition AmplitudesTable 2.2: Low-ly<strong>in</strong>g energy levels of the double-well potential, obta<strong>in</strong>ed by diagonalizationus<strong>in</strong>g level p = 18 effective action. The parameters used: k 2 = −1,k 4 = 12, L = 16, ∆ = 0.1, t = 0.05. The errors are estimated by comparisonwith the diagonalization results obta<strong>in</strong>ed from higher-order effective actions, f<strong>in</strong>erdiscretizations, larger space cutoffs, and lower values of the propagation time t.k E k |∆E k | δE k0 0.328826502590357561530(2) 7 × 10 −22 2 × 10 −211 1.41726810105965210733(23) 5 × 10 −21 4 × 10 −212 3.0819506284815341204(849) 3 × 10 −20 1 × 10 −203 5.019323060355788021(7990) 2 × 10 −19 4 × 10 −204 7.186203252338934478(3958) 5 × 10 −19 8 × 10 −205 9.54285734251209386(72421) 2 × 10 −18 2 × 10 −196 12.06403774639116375(04211) 4 × 10 −18 4 × 10 −197 14.7314279571006902(462590) 1 × 10 −17 7 × 10 −198 17.5310745155383834(413592) 3 × 10 −17 2 × 10 −189 20.4519281359123716(968554) 5 × 10 −17 3 × 10 −18numerical results with some known properties of the physical system. One suchproperty is density of states, def<strong>in</strong>ed formally asρ(E) =∞∑δ(E − E k ) , (2.30)k=0assum<strong>in</strong>g that the system has a discrete spectrum. This highly relevant physicalquantity can be directly calculated us<strong>in</strong>g the numerically obta<strong>in</strong>ed spectra. On theother hand, it can be also analytically calculated us<strong>in</strong>g semiclassical approximation.This approximation is valid at least <strong>in</strong> the high-energy region, and we can use itto assess the quality of our numerical results. In semiclassical approximation, thedensity of states <strong>in</strong> d spatial dimensions is calculated asρ sc (E) =∫ d⃗xd⃗pδ(E − H(⃗x, ⃗p)) . (2.31)(2π)dreplac<strong>in</strong>g the discrete spectrum with a cont<strong>in</strong>uous distribution of energy def<strong>in</strong>ed bythe classical Hamilton function H(⃗x, ⃗p). After <strong>in</strong>tegration over momenta, we obta<strong>in</strong>42


2. Diagonalization of Transition Amplitudesthe well known result [52]ρ sc (E) =( 1) d/2 ∫2π 2Γ (d/2)d⃗x Θ(E − V (⃗x)) (E − V (⃗x)) d/2−1 , (2.32)where Θ is the Heaviside step-function and Γ is the Gamma function. For the quarticanharmonic potential (2.29) <strong>in</strong> d = 1 the density of states can be expressed <strong>in</strong> termsof the complete elliptic <strong>in</strong>tegral of the first k<strong>in</strong>d K(k) = F(π/2, k) [63],ρ sc (E) =√ (√)2/π2 2 1(k2/4 2 + k 4 E/6) 1/4K 2 − k 2 /4√ . (2.33)k22 /4 + k 4 E/6In practical applications, especially <strong>in</strong> d = 1, it might be difficult to comparedirectly semiclassical approximation for density of states and numerically obta<strong>in</strong>edhistogram for ρ(E), s<strong>in</strong>ce energy levels are usually not degenerated, so the spectrumis very sparse. In order to have sufficient statistics for a reasonable histogram, onehas to use large value for b<strong>in</strong> size, and effectively the whole numerically availablespectrum is reduced to just a few b<strong>in</strong>s. For this reason, it is more <strong>in</strong>structive tostudy the cumulative density of states,n(E) =∫ EV m<strong>in</strong>dE ′ ρ(E ′ ) , (2.34)which counts the number of energy eigenstates smaller or equal to E. For quarticanharmonic oscillator the cumulative density of states is given by the above <strong>in</strong>tegralof the complete elliptic <strong>in</strong>tegral of the first k<strong>in</strong>d, and can be calculated numerically.Fig. 2.10 gives comparison of cumulative density of states calculated fromour numerical diagonalization results and semiclassical approximation n sc (E). Asexpected, the agreement is excellent up to very high values of energies, where numericaldiagonalization eventually fails due to the f<strong>in</strong>ite number of calculated energyeigenvalues and effects of discretization. Such behavior can be improved by us<strong>in</strong>gf<strong>in</strong>er discretization (smaller spac<strong>in</strong>g size), as illustrated by two different discretizationsteps for k 4 = 48, <strong>in</strong> Fig. 2.10. Such analysis can be used to assess the obta<strong>in</strong>edspectrum and determ<strong>in</strong>e the number of reliable energy eigenvalues. Typically wecan achieve up to 10 4 reliable energy eigenlevels with simulations on a s<strong>in</strong>gle CPU.In order to further demonstrate the applicability of the method, we also present43


2. Diagonalization of Transition Amplitudes300250200k 4 = 12, k 2 = -10, L / ∆ = 8192k 4 = 12, k 2 = -1, L / ∆ = 8192k 4 = 48, k 2 = 1, L / ∆ = 8192k 4 = 48, k 2 = 1, L / ∆ = 100n(E)1501005000 200 400 600 800 1000 1200 1400 1600 1800EFigure 2.10: Cumulative distribution of the density of numerically obta<strong>in</strong>ed energyeigenstates for the quartic anharmonic (k 2 = 1) and double-well potential (k 2 = −1),for t = 0.02, p = 21 and the follow<strong>in</strong>g values of diagonalization parameters: L = 10for k 4 = 12 and L = 8 for k 4 = 48. The discretization step is given on the graph bythe value of L/∆, top to bottom. Long-dashed l<strong>in</strong>es give correspond<strong>in</strong>g semiclassicalapproximations for the cumulative density of states.numerical results for the modified Pöschl-Teller modelV (x) = − χ22λ(λ − 1)cosh 2 χx , (2.35)which has only a f<strong>in</strong>ite set of discrete energy eigenlevels E k = −χ 2 (λ −1−k) 2 /2 for<strong>in</strong>teger k from the <strong>in</strong>terval 0 ≤ k ≤ λ − 1. Energy eigenvalues and eigenfunctions ofthis model are analytically known, and we will use them to further test our method.Effective actions to very high order are available also for this potential [60], andwe use them for numerical diagonalization of the evolution operator. Naturally, thediagonalization will give as many eigenvalues and eigenvectors as the size of thematrix S, but only the first few can be <strong>in</strong>terpreted as bound states of the potential,accord<strong>in</strong>g to the above condition 0 ≤ k ≤ λ − 1.Fig. 2.11(top) gives the analysis of errors <strong>in</strong> the ground energy due to the spacecutoff, while Fig. 2.11(bottom) gives the correspond<strong>in</strong>g analysis of L-errors for numericalcalculation of the energy level E 5 . As we can see, the behavior of errors is thesame as for the case of anharmonic oscillator, and we are aga<strong>in</strong> able to obta<strong>in</strong> high44


||2. Diagonalization of Transition Amplitudes| E 0(p) (∆, L, t) - E 0exact10 -5 110 -1010 -1510 -2010 -25p = 1p = 5p = 9p = 13p = 17p = 211 2 3 4 5 6L| E 5(p) (∆, L, t) - E 5exact10 -5 110 -1010 -1510 -2010 -25p = 1p = 5p = 9p = 13p = 17p = 211 2 3 4 5 6 7 8LFigure 2.11: Deviations |E (p)k(∆, L, t) − Eexact k | as a function of L for k = 0 (top)and k = 5 (bottom), for the modified Pöschl-Teller potential. Energy eigenvaluesare obta<strong>in</strong>ed us<strong>in</strong>g effective action levels p = 1, 5, 9, 13, 17, 21 and t = 0.1, with theparameters χ = 0.5, λ = 15.5, ∆ = 0.02.accuracy results. Fig. 2.12 gives the time dependence of errors <strong>in</strong> ground energyobta<strong>in</strong>ed by numerical diagonalization us<strong>in</strong>g different levels p of effective actions.The scal<strong>in</strong>g of errors proportional to t p is evident from the graph, as well as the discretizationerrors due to the f<strong>in</strong>ite discretization step ∆. To ensure that the effectivepotential is bounded from below, <strong>in</strong> this case we have to remove higher-order powersof discretized velocity δ from the effective potential near x = 0, s<strong>in</strong>ce such termshave non-vanish<strong>in</strong>g negative coefficients <strong>in</strong> the vic<strong>in</strong>ity of x = 0, due to a peculiarnature of the potential. In practical applications, one can use e.g. p = 1 effective45


|2. Diagonalization of Transition Amplitudes10 5 0.001 0.01 0.1| E 0(p) (∆, L, t) - E 0exact10 -5 110 -1010 -1510 -20∆ = 0.05∆ = 0.10∆ = 0.2010 -25Figure 2.12: Deviations |E (p)k(∆, L, t) − Eexact k | as a function of t for k = 0, for themodified Pöschl-Teller potential. Energy eigenvalues are obta<strong>in</strong>ed us<strong>in</strong>g effectiveaction levels p = 1, 3, 5, 7, 9, 11, 13 and L = 5, with the parameters χ = 0.5, λ = 15.5,∆ = 0.02. Dashed l<strong>in</strong>es <strong>in</strong> correspond to the discretization error (2.19).taction (which does not depend on δ) near x = 0. As can be seen, this does not affectthe obta<strong>in</strong>ed numerical results.Table 2.3(top) gives the obta<strong>in</strong>ed energy spectra for the modified Pöschl-Tellerpotential with the parameters χ = 0.5, λ = 15.5. If necessary, the precision ofobta<strong>in</strong>ed energy levels can be further <strong>in</strong>creased by appropriately chang<strong>in</strong>g the discretizationparameters. Contrary to the situation for anharmonic oscillator, whererelative error of numerically calculated low-ly<strong>in</strong>g energy levels did not change significantly,here we see that the <strong>in</strong>crease <strong>in</strong> the error is substantial. This is causedby the fact that this potential has only a small f<strong>in</strong>ite set of discrete bound states, soenergy levels k ∼ 10 correspond to the very top of the discrete spectrum. In practicalapplications such pathological situations are not encountered, but as we can see,even this can be dealt with by the proper choice of discretization parameters. Thequality of numerically calculated eigenfunctions is assessed <strong>in</strong> Table 2.3(bottom),where we give a symmetric matrix of scalar products 〈ψ k |ψlexact 〉 of numerically calculatedand analytic eigenfunctions. As we can see, the overlap between analyticand numeric eigenfunctions is excellent, and they are orthogonal with high precision,which is preserved even for higher energy levels. We have also verified that forparameters given <strong>in</strong> the caption of Table 2.3 and with the discretization step of theorder ∆ = 10 −3 eigenfunctions of all bound states can be accurately reproduced.46


2. Diagonalization of Transition AmplitudesTable 2.3: Top: Low-ly<strong>in</strong>g energy levels of the modified Pöschl-Teller potential,obta<strong>in</strong>ed by diagonalization us<strong>in</strong>g level p = 21 effective action. The parametersused: χ = 0.5, λ = 15.5, L = 5, ∆ = 0.02, t = 0.1. Bottom: Symmetric tableof scalar products 〈ψ k |ψlexact 〉 of numerically calculated and analytic eigenstates fork, l = 0, 1, 2, 3, 4.k E k E exactk|E k − E exactk | δE k0 −26.28125000000000000000000(174) −26.28125 2 × 10 −24 7 × 10 −261 −22.781250000000000000000(28812) −22.78125 3 × 10 −22 2 × 10 −232 −19.53124999999999999999(736443) −19.53125 3 × 10 −21 2 × 10 −223 −16.5312499999999999999(6571136) −16.53125 4 × 10 −20 2 × 10 −214 −13.7812499999999999(8195897101) −13.78125 2 × 10 −17 2 × 10 −185 −11.28124999999999(398393103608) −11.28125 6 × 10 −15 6 × 10 −166 −9.03124999999(8602255352218206) −9.03125 2 × 10 −12 2 × 10 −137 −7.031249999(773547728177905754) −7.03125 3 × 10 −10 4 × 10 −118 −5.2812499(74811672590174261082) −5.28125 3 × 10 −8 5 × 10 −90 1 2 3 40 1 − 4 · 10 −12 1.2 · 10 −13 2.8 · 10 −6 7.4 · 10 −14 3.0 · 10 −71 1.2 · 10 −13 1 − 1 · 10 −11 2.1 · 10 −13 4.5 · 10 −6 4.6 · 10 −142 2.8 · 10 −6 2.1 · 10 −13 1 − 2 · 10 −11 1.3 · 10 −13 5.9 · 10 −63 7.4 · 10 −14 4.5 · 10 −6 1.3 · 10 −13 1 − 4 · 10 −11 3.1 · 10 −134 3.0 · 10 −7 4.6 · 10 −14 5.9 · 10 −6 3.1 · 10 −13 1 − 5 · 10 −112.5 Numerical results for two-dimensional modelsIn this section we illustrate the application of the numerical method based on thediagonalization of transition amplitudes on two models <strong>in</strong> d = 2 spatial dimensions.The first model is the anharmonic oscillatorV (x, y) = k 22 (x2 + y 2 ) + k 424 (x2 + y 2 ) 2 , (2.36)which is used for a description of the trapp<strong>in</strong>g potential used <strong>in</strong> a recent experimentwith fast-rotat<strong>in</strong>g Bose-E<strong>in</strong>ste<strong>in</strong> condensate of 87 Rb atoms [12, 64, 65].The graphs <strong>in</strong> Fig. 2.13 are calculated for k 2 = 0, when the potential is reducedto a pure quartic <strong>in</strong>teraction. The analysis of errors is very similar as <strong>in</strong> the onedimensionalcases we studied <strong>in</strong> the previous section. The dependence of groundenergy errors on the space cutoff L is shown <strong>in</strong> Fig. 2.13(top), and we see the47


2. Diagonalization of Transition Amplitudes10 2 0 0.5 1 1.5 2 2.5 3|| E 0(p) (∆, L, t) - E 0exact10 -2 110 -410 -610 -810 -1010 -12p = 1p = 2p = 4p = 6p = 8p = 10p = 12p = 1410 -2 1L10 -4| E 0(p) (∆, L, t) - E 0exact|10 -610 -810 -1810 -1210 -1410 -1610 -180.020.04t0.060.08p = 1p = 2p = 4p = 6p = 8p = 100.10.12Figure 2.13: Deviations from the ground energy |E (p)0 (∆, L, t) − E 0 | as a function ofthe space cutoff L (top) and as a function of the time t (bottom) for the potential(2.36) for k 2 = 0 and k 4 = 24. The discretization parameters are ∆ = 0.2, t = 0.1on the top graph, and L = 3 on the bottom graph. Deviations are calculated us<strong>in</strong>gthe ground energy E 0 = 1.47714975357799(4) obta<strong>in</strong>ed with p = 21 effective action.The dashed l<strong>in</strong>e <strong>in</strong> the bottom graph corresponds to the discretization error (2.19).usual saturation of errors for sufficiently large values of L. The saturated valuerapidly decreases (by several orders of magnitude) as we <strong>in</strong>crease the level p of theeffective action used to calculate space-discretized matrix of the evolution operator.Fig. 2.13(bottom) shows the time dependence of ground energy errors, which are48


2. Diagonalization of Transition AmplitudesTable 2.4: Low-ly<strong>in</strong>g energy levels of a d = 2 anharmonic potential (2.36), obta<strong>in</strong>edus<strong>in</strong>g the level p = 21 effective action. The discretization parameters are L = 14,∆ = 0.14, and t = 0.2.k E k , k 2 = −0.1025, k 4 = k exp4 E k , k 2 = −0.1025, k 4 = 10 3 k exp40 -1.1279858856602 1.12872978314351 -1.1169327267787 2.61613484978342 -1.1169327267787 2.61613484978343 -1.0842518375067 4.34765152798104 -1.0842518374840 4.34765152798125 -1.0311383813261 4.65284518520136 -1.0311383813261 6.27045529036717 -0.95910186300510 6.27045529036718 -0.95910186300478 6.75898824914119 -0.86968170695135 6.7589882491412found to fully agree with the scal<strong>in</strong>g law t p for sufficiently f<strong>in</strong>e discretization. Aga<strong>in</strong>,the discretization errors conform to the universal dependence given <strong>in</strong> Eq. (2.19).Here there is an additional factor of 2 <strong>in</strong> the cosh term due to the dimensionality ofthe system.Table 2.4 gives the numerically obta<strong>in</strong>ed energy eigenvalues for different sets ofparameters of the potential (2.36). Motivated by the values of the experimentalparameters [12, 64], we <strong>in</strong>troduce and use the constant k exp4 = 1.95×10 −3 . From theanalysis of discretization errors and errors related to the use of a chosen effectiveaction level p, we can estimate the errors <strong>in</strong> found energy eigenvalues to be of theorder 10 −15 . The results <strong>in</strong> the Table 2.4 are obta<strong>in</strong>ed by numerical diagonalizationbased on the C SPEEDUP code [60] and the use of the LAPACK [62] library. Theestimated error <strong>in</strong> energy eigenvalues is smaller than the (relative) error which canbe achieved <strong>in</strong> typical C simulations, which is of the order 10 −14 . This is easilyverified, s<strong>in</strong>ce for several different values of discretization parameters we get thesame stable results shown <strong>in</strong> the table. Therefore, this table gives certa<strong>in</strong> digits<strong>in</strong> all energy eigenvalues, and the error can be cited as implicit (half of the lastdigit). This is good example for practical applications, where we have managedto elim<strong>in</strong>ate all types of errors below the limit that can be seen due to <strong>in</strong>herentnumerical errors of computer simulation. However, if such complete elim<strong>in</strong>ation oferrors is not possible due to the limitations <strong>in</strong> computer memory or computation49


2. Diagonalization of Transition Amplitudestime, the analysis of errors presented <strong>in</strong> Fig. 2.13 allows us to reliably estimatenumerical errors <strong>in</strong> energy eigenvalues.Fig. 2.14 shows the numerically obta<strong>in</strong>ed ground state for this two-dimensionalpotential for the case of k 2 < 0. The ground state has the expected Mexican-hatshape. The figure gives a three-dimensional plot of the ground state on the left,and the correspond<strong>in</strong>g density plot on the right, with values of the wave functionmapped to colors. Fig. 2.15 gives density plots of k = 1, 2, 3, 4 eigenfunctions forthe same values of parameters. The discretization is sufficiently f<strong>in</strong>e (∆ = 0.25) sothat all features of calculated eigenfunctions are clearly visible.As <strong>in</strong> the one-dimensional case, we will calculate the density of states ρ sc (E)<strong>in</strong> semiclassical approximation, and use it as a criterion for the reliability of highenergyeigenstates. In d = 2, the density of states is given by a simple formulaρ sc (E) = 1 ∫ ∫2πdxdy Θ(E − V (x, y)) . (2.37)For the quartic anharmonic potential (2.36) the density of states can be analyticallycalculatedρ sc (E) = − 3 k 2k 4+√9 k 2 2k 2 4+ 6Ek 4. (2.38)Fig. 2.16(top) shows the comparison of semiclassical approximation for the densityof states, and the histogram for numerically obta<strong>in</strong>ed energy eigenvalues of the-15|ψ|0.020.010.020.010y-10-5050.020.010010-10-50x510-10 -5 05y1015-15 -10 -5 0 5 10 15xFigure 2.14: Ground state (as 3-D plot on the left, and as a density plot on theright) of a d = 2 anharmonic potential (2.36) obta<strong>in</strong>ed us<strong>in</strong>g p = 21 effective action.The parameters are k 2 = −0.1025, k 4 = k exp4 , L = 20, ∆ = 0.25, t = 0.2.50


2. Diagonalization of Transition Amplitudesy-15-10-5050.040.020-0.02-0.04y-15-10-5050.040.020-0.02-0.04101015-15 -10 -5 0 5 10 15x15-15 -10 -5 0 5 10 15xy-15-10-5050.040.020-0.02-0.04y-15-10-5050.040.020-0.02-0.04101015-15 -10 -5 0 5 10 15x15-15 -10 -5 0 5 10 15xFigure 2.15: Density plots of level k = 1, 2, 3, 4 eigenstates of a d = 2 anharmonicpotential (2.36) obta<strong>in</strong>ed us<strong>in</strong>g p = 21 effective action. The parameters are k 2 =−0.1025, k 4 = k exp4 , L = 20, ∆ = 0.25, t = 0.2.potential (2.36). Due to the high degeneracy of energy eigenstates <strong>in</strong> d = 2, thehistogram of numerically found energy levels conta<strong>in</strong>s enough statistics over thewhole region of energies, and therefore can be used for assessment of the quality ofnumerical spectra. As we see, the agreement is better and better when we use f<strong>in</strong>erspace discretization. Depend<strong>in</strong>g on the needed number of energy levels and maximalvalue of the energy considered to be relevant for the calculation we can chooseappropriate values of discretization parameters that will provide reliable numericalresults up to desired energy value. For example, for the choice of discretizationparameters L = 14, ∆ = 0.14, we can reliably use energy levels up to E ≈ 120.Fig. 2.16(bottom) shows the comparison of cumulative density of states n(E)calculated for numerically obta<strong>in</strong>ed results and <strong>in</strong> semiclassical approximation, by51


2. Diagonalization of Transition Amplitudes250200L = 14.4, ∆ = 0.14L = 14.4, ∆ = 0.24L = 10, ∆ = 0.25L = 5, ∆ = 0.251501005003500030000250000 50 100 150 200 250L = 14.4, ∆ = 0.14L = 14.4, ∆ = 0.24L = 10, ∆ = 0.25L = 5, ∆ = 0.25200001500010000500000 50 100 150 200 250 300Figure 2.16: Distribution of the density of numerically obta<strong>in</strong>ed energy eigenstates(top) and cumulative distribution of the density of numerically obta<strong>in</strong>ed energyeigenstates (bottom) for a d = 2 anharmonic potential (2.36), calculated with thelevel p = 21 effective action. The parameters are k 2 = 1, k 4 = k exp4 , t = 0.2, whilediscretization parameters are given on the graph, correspond<strong>in</strong>g to the curves topto bottom. Long-dashed l<strong>in</strong>es on both graphs give the correspond<strong>in</strong>g semiclassicalapproximations.<strong>in</strong>tegrat<strong>in</strong>g the expression (2.38), which can be calculated analytically. The comparisonof numerical and semiclassical cumulative density of states <strong>in</strong> Fig. 2.16(bottom)verifies our conclusions from Fig. 2.16(top), and aga<strong>in</strong> sets the same limit of reliableenergy levels for chosen discretization parameters.52


2. Diagonalization of Transition AmplitudesThe second two-dimensional model we have studied numerically is a sextic anharmonicoscillator,V (x, y) = V x (x) + V y (y) + V xy (x − y) , (2.39)where V i (x) = V i0 (a i x 2 + b i x 4 + c i x 6 ). The values of the coefficients used are given<strong>in</strong> Table 2.5. The study of this potential is motivated by Ref. [66], where it has beenused to <strong>in</strong>vestigate the transition from regular to chaotic classical motion. Fig. 2.17shows the numerically obta<strong>in</strong>ed ground state for this two-dimensional potential, asa three-dimensional plot on the left, and as a density plot on the right. Fig. 2.18gives density plots of k = 1, 3, 7, 8 eigenfunctions for the same values of parameters.The discretization is sufficiently f<strong>in</strong>e (∆ = 0.04) so that we can resolve all details <strong>in</strong>the presented eigenstates.Table 2.5: Parameters of the sextic potential (2.39).i V i0 a i b i c ix 100 1.56 -0.61 0.32y 100 0.69 -0.12 0.03xy 100 -1.00 0.25 0.08-20.075|ψ|0.0750.050.0250-1x01-10y10.0750.050.02502y-1012-2 -1 0 1 2x0.050.0250Figure 2.17: Ground state (as 3-D plot on the top, and as a density plot on thebottom) of a sextic anharmonic potential, obta<strong>in</strong>ed by diagonalization us<strong>in</strong>g thelevel p = 21 effective action. The parameters of the potential are given <strong>in</strong> the text.The diagonalization parameters: L = 4, ∆ = 0.04, t = 0.01.53


2. Diagonalization of Transition Amplitudesnential growth <strong>in</strong> the size of analytic expressions for the effective potential with the<strong>in</strong>crease of the level p, as discussed <strong>in</strong> Ref. [44]. Therefore, the required CPU timefor construction of the matrix to be diagonalized <strong>in</strong> the presented approach growsexponentially with the level p, while <strong>in</strong> other methods the construction of such amatrix does not require a significant amount of time. However, the time for exactdiagonalization far outweighs the time needed for construction of even large matriceswith moderate levels p of the order 10-20. The significant benefit of practicallyelim<strong>in</strong>at<strong>in</strong>g errors associated with the time of propagation therefore fully justifiesthe use of the effective action approach. Of course, <strong>in</strong> practical applications one hasto study the complexity of the algorithm and to choose the optimal level p whichwill sufficiently reduce the errors, while keep<strong>in</strong>g the complexity of the calculation onthe acceptable level.2.6 Conclusions and outlookIn this Chapter, we have dealt with the thorough understand<strong>in</strong>g and optimizationof the method of the calculation of the properties of quantum systems based on thediagonalization of transition amplitudes, previously <strong>in</strong>troduced <strong>in</strong> Ref. [9]. First,we have focused on analyz<strong>in</strong>g the errors associated with real-space discretizationand f<strong>in</strong>ite size effects. In particular, we have shown that with<strong>in</strong> this calculationscheme spatial discretization leads to a universal and non-perturbatively small discretizationerror. This highly outperforms the usual polynomial behavior of errors<strong>in</strong> approaches based on the diagonalization of space-discretized Hamiltonians. Akey problem <strong>in</strong> practical applications of this approach - accurate calculation of transitionamplitudes, matrix elements of the space-discretized evolution operator, hasbeen resolved us<strong>in</strong>g recently <strong>in</strong>troduced effective action approach [44], which givessystematic short-time expansion of the evolution operator.The derived analytical estimates for all types of errors, <strong>in</strong>clud<strong>in</strong>g errors due to theapproximative calculation of transition amplitudes, provide us with a way to chooseoptimal discretization parameters and to reduce overall errors <strong>in</strong> energy eigenvaluesand eigenstates for many orders of magnitude, as was demonstrated for several oneandtwo-dimensional models. We have shown that numerical diagonalization of thespace-discretized evolution operator can be successfully applied for studies of many<strong>in</strong>terest<strong>in</strong>g lower dimensional models. The approach allows exact numeric calculationof a large number of energy eigenvalues and eigenstates of the system. Due to55


2. Diagonalization of Transition Amplitudesthe superior behavior of discretization and other errors <strong>in</strong> this method comparedto methods based on diagonalization of the discretized Hamilton operator and relatedmethods, the presented approach is a method of choice for numerical studiesof lower-dimensional physical systems. An application of the approach for numerical<strong>in</strong>vestigation of properties of fast rotat<strong>in</strong>g Bose-E<strong>in</strong>ste<strong>in</strong> condensates is given<strong>in</strong> Chapter 3. Further <strong>in</strong>terest<strong>in</strong>g l<strong>in</strong>e of research would be to comb<strong>in</strong>e the presentmethod with the density matrix renormalization group (DMRG) approach [48, 67].56


Chapter 3Thermodynamics of a rotat<strong>in</strong>g ideal BECThe behavior of a BEC under rotation is essential for understand<strong>in</strong>g many fundamentalphenomena [68, 7, 19]. The response of a quantum fluid to rotation representsone of the sem<strong>in</strong>al hallmarks of superfluidity characterized by a nucleationof vortices with a quantized circulation. In the past, quantum vortices were studiedexperimentally <strong>in</strong> the superfluid helium and <strong>in</strong> type-II superconductors. An importantadvantage of the cold atomic gases for study<strong>in</strong>g vortices is that the typical sizeof a vortex core is 3 orders of magnitude larger than <strong>in</strong> the superfluid Helium, and,even more importantly, it is large enough to be observed optically [69, 7]. Earlyexperiments with rotat<strong>in</strong>g cold gases explored different regimes: for slow rotation asmall number of vortices was observed [70, 69], while the triangular Abrikosov lattice[71] composed of about 100 vortices was detected <strong>in</strong> the case of a higher rotationfrequency [72]. Eventually, <strong>in</strong> the case of a very rapid rotation it is expected thatbosons would go from the superfluid phase <strong>in</strong>to a strongly correlated phase that isrelated to the quantum Hall physics [73].The equivalence of the rotation and the motion of a charged particle <strong>in</strong> a magneticfield can be easily understood by consider<strong>in</strong>g the Hamiltonian of a s<strong>in</strong>gle particle <strong>in</strong>a harmonic trap <strong>in</strong> a rotat<strong>in</strong>g reference frame with the rotation frequency ⃗ Ω = Ω⃗e z[68]:Ĥ rot = Ĥ − Ω ⃗ · ˆ⃗ L= 12M (ˆp2 x + ˆp2 y + ˆp2 z ) + 1 2 Mω2 (ˆx 2 + ŷ 2 + λ 2 zẑ2 ) − Ω⃗e z · (ˆ⃗r × ˆ⃗p)= 12M (ˆp x + MΩŷ) 2 + 12M (ˆp y − MΩˆx) 2+ 1 2 M(ω2 − Ω 2 )(ˆx 2 + ŷ 2 ) + 12M ˆp2 z + 1 2 Mλ2 z ω2 ẑ 2 , (3.1)where ˆ⃗ L is the angular momentum and all the variables <strong>in</strong> the last expression aregiven <strong>in</strong> the rotat<strong>in</strong>g reference frame. From Eq. (3.1), we see that <strong>in</strong> the limit Ω → ω,the Hamiltonian becomes formally equivalent to the one describ<strong>in</strong>g particles <strong>in</strong> the57


magnetic field ⃗ B = 2M ⃗ Ω.3. Rotat<strong>in</strong>g ideal BECHowever, once harmonically trapped Bose-E<strong>in</strong>ste<strong>in</strong> condensate is rotated critically,i.e. the rotation frequency becomes so large that it fully compensates theradially conf<strong>in</strong><strong>in</strong>g harmonic trapp<strong>in</strong>g, the system turns out to be radially no longerconf<strong>in</strong>ed. In the absence of additional potential terms, the condensate would start toexpand <strong>in</strong> directions perpendicular to the rotation axis. For an overcritical rotation,this expansion would even be accelerated by the presence of a residual centrifugalforce. This poses a severe problem to the experimental achievement of strongly correlatedbosonic states. In order to reach experimentally this delicate regime, Fettersuggested <strong>in</strong> Ref. [74] a small quartic term to be added to the harmonic trap potential,which would provide a conf<strong>in</strong>ement <strong>in</strong> the radial direction <strong>in</strong> the case of acritical and overcritical rotation.The proposal has been experimentally realized <strong>in</strong> Paris by the Dalibard group fora BEC of 87 Rb atoms [12, 7], by superimpos<strong>in</strong>g to the magnetic trap an additionalGaussian laser beam propagat<strong>in</strong>g <strong>in</strong> the z-direction,U(x, y) = U 0 e −2(x2 +y 2 )w 2 . (3.2)In the previous equation, w denotes the laser beam waist and U 0 is the <strong>in</strong>tensity ofthe beam. With<strong>in</strong> the laser beam waist, the potential can be approximated by:U(x, y) ≈ U 0 − 2 U 0x 2 + y 2w 2 + 2 U 0(x 2 + y 2 ) 2w 4 , (3.3)and this is how the quartic term is brought about. An additional laser beam thatcreates a small anisotropic potential <strong>in</strong> the x − y plane is used for stirr<strong>in</strong>g the pureBEC at the rotation frequency Ω. After the angular momentum has been <strong>in</strong>troduced<strong>in</strong>to the system, the rotation is stopped and the condensate is allowed to equilibrate.Tak<strong>in</strong>g <strong>in</strong>to account Eqs. (3.1) and (3.3), the result<strong>in</strong>g axially-symmetric trap witha small quartic anharmonicity <strong>in</strong> the x − y plane, seen by <strong>in</strong>dividual atoms, has theformV BEC = M 2 (ω2 − Ω 2 )(x 2 + y 2 ) + M 2 ω2 z z2 + κ 4 (x2 + y 2 ) 2 , (3.4)with the trap frequencies ω = 2π × 64.8 Hz, ω z = 2π × 11.0 Hz, and the trapanharmonicity κ = κ BEC = 2.6 × 10 −11 Jm −4 . The rotation frequency Ω, measured<strong>in</strong> units of ω and expressed by the ratio η = Ω/ω, represents the tunable controlparameter, which could be experimentally varied <strong>in</strong> the range between 0 and 1.04.58


3. Rotat<strong>in</strong>g ideal BEC Figure 3.1: Images of a rotat<strong>in</strong>g BEC along the rotation direction for differentrotation frequencies Ω/2π. The l<strong>in</strong>ear size of each image is 306 µm. The results aretaken from Ref. [12].To probe the system, the TOF absorption imag<strong>in</strong>g is performed. Typical resultsof the measurements for different rotation frequencies are presented <strong>in</strong> Fig. 3.1. Itis obvious that up to Ω = 2π × 68 Hz the radius of a trapped cloud <strong>in</strong>creaseswith <strong>in</strong>crease <strong>in</strong> Ω. For Ω = 2π × 66 Hz and Ω = 2π × 67 Hz the radial densityprofile of the cloud follows the Mexican-hat shape of the potential, however numberof observed vortices <strong>in</strong> this case is smaller than expected for such a large rotationfrequency. Several possible schemes are discussed as a possible explanation of theobserved features.In order to contribute to the understand<strong>in</strong>g of the experimental results, we studythe BEC phase transition of an ideal Bose gas <strong>in</strong> the trapp<strong>in</strong>g potential (3.4), modifiedby the presence of a quartic term and a rotation with respect to the commonharmonic trap (1.6). Depend<strong>in</strong>g on the value of the rotation frequency, the shape ofthe potential changes from convex with a s<strong>in</strong>gle m<strong>in</strong>imum to the Mexican-hat shape,which significantly <strong>in</strong>fluences the properties of a condensate. To study these effects,<strong>in</strong> the rest of this Chapter we calculate the condensation temperature, condensatefraction and density profiles of the cloud <strong>in</strong> the trap (3.4) and also simulate a freeexpansion of the condensate, correspond<strong>in</strong>g to the TOF imag<strong>in</strong>g.As long as we approximately describe the system with the ideal Bose gas, allof its many-body properties <strong>in</strong> the grand-canonical ensemble can be derived purelyfrom s<strong>in</strong>gle-particle states. When consider<strong>in</strong>g the thermodynamic limit, usually thesemiclassical approximation is applied, where the s<strong>in</strong>gle-particle ground state E 0is reta<strong>in</strong>ed and treated quantum mechanically, while all other excited states aretreated as a cont<strong>in</strong>uum [19, 65]. The validity of the semiclassical approximationcan be rigorously justified <strong>in</strong> some regimes of the relevant parameters. As brieflymentioned <strong>in</strong> Chapter 1, the semiclassical approximation is justified <strong>in</strong> the hightemperaturelimit, when the thermal energy is larger than the typical spac<strong>in</strong>g ofenergy levels, k B T ≥ (E n+1 − E n ). Also, it rema<strong>in</strong>s reasonable good irrespective59


3. Rotat<strong>in</strong>g ideal BECof the rotation frequency Ω once the total particle number N is large enough andthe trap anharmonicity κ is small enough. The latter condition implies that theunderly<strong>in</strong>g potential (3.4) has a small curvature around its m<strong>in</strong>imum, and hencethe correspond<strong>in</strong>g density of energy levels is sufficiently high. However, <strong>in</strong> thiscontext the question arises how accurate the semiclassical approximation is, forwhich system parameters it is not anymore sufficient for a precise description ofBEC phenomena, as well as when it f<strong>in</strong>ally breaks down, requir<strong>in</strong>g a full quantummechanicaltreatment of the system.In order to analyze the problem more quantitatively, it is mandatory to determ<strong>in</strong>ethe s<strong>in</strong>gle-particle energy eigenvalues and eigenfunctions fully quantum mechanically.In this Chapter we show how the exact diagonalization of a time-evolutionoperator, presented <strong>in</strong> Chapter 2, is applied for study<strong>in</strong>g both global and local propertiesof fast-rotat<strong>in</strong>g Bose-E<strong>in</strong>ste<strong>in</strong> condensates. To this end we proceed as follows:first we calculate a large number of energy eigenvalues and eigenfunctions for the anharmonicpotential (3.4). Afterwards, we discuss how a f<strong>in</strong>ite number of numericallyavailable energy eigenvalues affects the results and how they can be improved by<strong>in</strong>troduc<strong>in</strong>g systematic semiclassical corrections. On the basis of this precise numericals<strong>in</strong>gle-particle <strong>in</strong>formation, we study global properties of a rotat<strong>in</strong>g condensate.F<strong>in</strong>ally, we calculate local properties of the condensate, such as density profiles andTOF absorption pictures.To beg<strong>in</strong> with, we rewrite Eq. (1.3) for the total number of particles <strong>in</strong> a moreconvenient form <strong>in</strong> terms of the s<strong>in</strong>gle-particle partition function Z 1 (β), def<strong>in</strong>ed asZ 1 (β) =∞∑e −βEn . (3.5)n=0To do this, we s<strong>in</strong>gle out the contribution of the ground state and use the Taylor’sexpansion 1/(1−x) = ∑ ∞n=0 xn (valid for |x| < 1), to derive the follow<strong>in</strong>g expression:N =∞∑n=01e β(En−µ) − 1 = B 0(µ, T) +∞∑n=1 j=1∞∑e −jβ(En−µ) . (3.6)60


3. Rotat<strong>in</strong>g ideal BECBy rearrang<strong>in</strong>g the summation order <strong>in</strong> the previous equation, we f<strong>in</strong>ally obta<strong>in</strong>:N = B 0 (µ, T) += B 0 (µ, T) +∞∑ ∑ ∞e jβµ e −jβEn ,j=1n=1∞∑e ( jβµ Z 1 (jβ) − e ) −jβE 0, (3.7)j=1where the summation on the right-hand side of the previous equation corresponds tothe cumulant expansion. In order to avoid any double-count<strong>in</strong>g, we have subtractedthe contribution of the ground state with<strong>in</strong> the s<strong>in</strong>gle-particle partition functionbecause a possible macroscopic occupation of the ground state is separately taken<strong>in</strong>to account.The BEC phase transition is achieved only <strong>in</strong> the thermodynamic limit of an<strong>in</strong>f<strong>in</strong>ite number of atoms, thus mak<strong>in</strong>g numerical studies of the condensation <strong>in</strong>creas<strong>in</strong>glydifficult. Usually, the problem is solved by fix<strong>in</strong>g the chemical potential µat the low temperatures of the condensate phase to the ground-state energy, i.e. bysett<strong>in</strong>g µ = E 0 . This requires that the ground state is treated separately by associat<strong>in</strong>ga macroscopic value N 0 to the ground-state occupation number B 0 (µ, T), forT < T c . Thus, from Eq. (3.7), the total number of particles <strong>in</strong> the condensate phasefollows to be:∞∑ (N = N 0 + ejβE 0Z 1 (jβ) − 1 ) . (3.8)j=1The equation (3.8) yields the temperature dependence of N 0 . With<strong>in</strong> the gas phase,where the macroscopic occupation of the ground state vanishes, i.e. we have N 0 =0, Eq. (3.7) determ<strong>in</strong>es the temperature dependence of the chemical potential µ.Therefore, the value of β c = 1/k B T c , which characterizes the boundary betweenboth phases, follows from Eq. (3.8) by sett<strong>in</strong>g N 0 = 0 and µ = E 0 :N =∞∑ [ejβ cE 0Z 1 (jβ c ) − 1 ] . (3.9)j=1We conclude that, for a given number N of ideal bosons, the condensation temperaturecan be exactly calculated only if both the s<strong>in</strong>gle-particle ground-state energyE 0 and the full temperature dependence of the one-particle partition function (3.5)are known.61


3. Rotat<strong>in</strong>g ideal BEC3.1 Numerical calculation of energy eigenvalues and eigenstatesA very efficient method for calculat<strong>in</strong>g properties of few-body quantum systems,that we use, is the direct diagonalization of the space-discretized propagator <strong>in</strong>imag<strong>in</strong>ary time. The approach is able to give very accurate energy eigenvalues evenfor moderate values of the propagation time t of the order 0.1, as shown <strong>in</strong> detail <strong>in</strong>Chapter 2. Note that throughout this Chapter we use dimensionless units, <strong>in</strong> whichall energies are expressed <strong>in</strong> terms of ω, while the length unit is the correspond<strong>in</strong>gharmonic oscillator length √ /Mω.Table 3.1 presents the first several energy eigenvalues for the two-dimensional(x − y) part of the BEC potential (3.4) for the non-rotat<strong>in</strong>g case (η = 0), as well asfor the critically-rotat<strong>in</strong>g condensate (η = 1). The table on the left gives the energyspectrum of the potential with the anharmonicity κ = κ BEC used <strong>in</strong> the experiment[12], while the right table shows the spectrum for the much larger anharmonicityκ = 10 3 κ BEC . The degeneracies of numerically obta<strong>in</strong>ed eigenstates <strong>in</strong> all casesTable 3.1: Lowest energy levels of the xy-part of the BEC potential (3.4) for nonrotat<strong>in</strong>g(η = 0) and critically rotat<strong>in</strong>g (η = 1) condensate with the quartic anharmonicityκ = κ BEC (left) and κ = 10 3 κ BEC (right). They are obta<strong>in</strong>ed by us<strong>in</strong>g levelp = 21 effective action with the discretization parameters of Table 3.3. The spac<strong>in</strong>g∆ was always chosen so that L/∆ = 100, and the propagation time was t = 0.2 forκ = κ BEC and t = 0.05 for κ = 10 3 κ BEC . Errors are given by the precision of thelast digit, typically 10 −12 to 10 −13 , and are estimated by compar<strong>in</strong>g the numericalresults obta<strong>in</strong>ed with different discretization parameters.E n /ω, κ = κ BECn η = 0 η = 10 1.0009731351803 0.11626671641341 2.0029165834022 0.26746899689052 2.0029165834022 0.26746899689053 3.0058275442161 0.44269273752694 3.0058275442161 0.44269273752705 3.0067964582067 0.47252757249416 4.0097032385903 0.63681788049837 4.0097032385903 0.63681788049848 4.0116368851078 0.68481424703569 4.0116368851078 0.6848142470357E n /ω, κ = 10 3 κ BECn η = 0 η = 10 1.468486725893 1.1626671641341 3.213056378201 2.6746899689052 3.213056378201 2.6746899689053 5.163819069871 4.4269273752694 5.163819069871 4.4269273752705 5.406908088225 4.7252757249416 7.282930987460 6.3681788049827 7.282930987460 6.3681788049828 7.690584058915 6.8481424703579 7.690584058915 6.84814247035762


3. Rotat<strong>in</strong>g ideal BECTable 3.2: Lowest energy levels of the xy-part of the BEC potential (3.4) for overcriticallyrotat<strong>in</strong>g (η = 1.04) condensate with the quartic anharmonicity κ = κ BECand κ = 10 3 κ BEC accord<strong>in</strong>g to the same numerical procedure as <strong>in</strong> Table 3.1.E n /ω, η = 1.04n κ BEC 10 3 κ BEC0 -0.6617041825660 1.1356938262061 -0.6465857464220 2.6281299037902 -0.6465857464220 2.6281299037903 -0.6032113415949 4.3638766339294 -0.6032113415948 4.3638766339295 -0.5349860004310 4.6676535829636 -0.5349860004309 6.2904447340077 -0.4451224795419 6.2904447340078 -0.4451224795419 6.7772107731699 -0.3362724309903 6.777210773169correspond to the expected structure of the spectrum, which can be deduced fromthe symmetry of the problem. In addition to this, the <strong>in</strong>terest<strong>in</strong>g case of criticalrotation (η = 1) allows a further verification of the numerical results. To this endwe recall that the energy eigenvalues of a pure quartic oscillator, to which V BECreduces <strong>in</strong> this case, are proportional to κ 1/3 due to a spatial rescal<strong>in</strong>g <strong>in</strong> the underly<strong>in</strong>gSchröd<strong>in</strong>ger equation. Therefore, we expect that the energy eigenvaluesfor κ = 10 3 κ BEC are precisely 10 times larger than the correspond<strong>in</strong>g eigenvaluesfor κ = κ BEC . Compar<strong>in</strong>g the rightmost columns <strong>in</strong> Table 3.1 we see exactly thisscal<strong>in</strong>g. This demonstrates conclusively that the presented method can be successfullyapplied also <strong>in</strong> this deeply non-perturbative parameter regime. Furthermore,Table 3.2 gives the energy spectrum of an over-critically rotat<strong>in</strong>g (η = 1.04) condensate,illustrat<strong>in</strong>g that the same approach can be used <strong>in</strong> this delicate regime aswell.With these results s<strong>in</strong>gle-particle partition functions Z 1 (β) can now be calculatedaccord<strong>in</strong>g to Eq. (3.5). This is especially suitable for the low-temperature regime,when higher energy levels give a negligible contribution. Although the above describedapproach is able to accurately give several thousands of energy eigenvalues,their number is always necessarily limited. This is easily seen from Fig. 3.2, wherewe compare the density of states for a critically rotat<strong>in</strong>g condensate with the corre-63


3. Rotat<strong>in</strong>g ideal BEC350300SC approx.L = 22.3, L/∆ = 100ρ(E)2502001501005000 20 40 60 80 100 120 140E / − hω ⊥Figure 3.2: Numerically calculated density of states for xy-part of the BEC potentialwith κ = κ BEC for a critically rotat<strong>in</strong>g condensate, obta<strong>in</strong>ed by us<strong>in</strong>g level p = 21effective action. The discretization parameters are L = 22.3, L/∆ = 100, andt = 0.2. The dashed l<strong>in</strong>e is the correspond<strong>in</strong>g semiclassical approximation for thedensity of states.spond<strong>in</strong>g semiclassical approximation for the density of states (2.38). This comparisonallows us to estimate the maximal reliable two-dimensional energy eigenvalueE max which can be obta<strong>in</strong>ed numerically for a given set of discretization parameters.For example, from Fig. 3.2 we can estimate E max ≈ 90 for η = 1 with the discretizationparameters L = 22.3, L/∆ = 100, and t = 0.2. Table 3.3 gives estimatesfor the maximal reliable energy eigenvalue for the anharmonicities κ = κ BEC andκ = 10 3 κ BEC for several values of rotation frequencies. These results are obta<strong>in</strong>edfrom numerical calculations us<strong>in</strong>g the SPEEDUP codes [60]. This table gives alsoan overview over those discretization parameters which were used for a numericaldiagonalization of the BEC potential (3.4) <strong>in</strong> order to calculate both global and localproperties of the condensate throughout this Chapter.In the low-temperature limit the f<strong>in</strong>iteness of the number of known energy eigenstatesdoes not present a problem. In fact, a precise knowledge of a large number ofenergy eigenvalues makes this approach a preferred method for a numerically exacttreatment of low-temperature phenomena. On the other hand, the high-temperatureregime, where thermal contributions of higher energy states play a significant role,is not treatable <strong>in</strong> the same way. This regime is usually not relevant for studies of64


3. Rotat<strong>in</strong>g ideal BECTable 3.3: Maximal reliable numerically calculated energy eigenvalue E max of thexy-part of the BEC potential (3.4) for different values of η = Ω/ω, estimated fromcompar<strong>in</strong>g the numerically obta<strong>in</strong>ed density of states ρ(E) with the semiclassicalapproximation. The numerical diagonalization was done us<strong>in</strong>g level p = 21 effectiveaction. The spac<strong>in</strong>g ∆ was always chosen so that L/∆ = 100, and the propagationtime was t = 0.2 for κ = κ BEC and t = 0.05 for κ = 10 3 κ BEC . The total number ofreliable energy eigenstates is <strong>in</strong> all cases of the order of 10 4 .κ = κ BEC κ = 10 3 κ BECη E max /ω L E max /ω L0.0 140 14.2 190 3.900.2 140 14.4 190 3.900.4 140 15.0 180 3.910.6 140 16.3 180 3.920.8 130 18.6 180 3.941.0 90 22.3 170 3.961.04 90 23.2 170 3.96BEC experiments, but we consider it for the sake of completeness. When the temperatureis sufficiently high, so that effects of higher energy eigenstates cannot beneglected, the <strong>in</strong>verse temperature β becomes a small parameter. Thus, it becomespossible to calculate numerically the s<strong>in</strong>gle-particle partition function as a sum ofdiagonal amplitudes, i.e.Z 1 (β) = Tr e −βĤ ≈ ∑ jA(j∆,j∆; β)∆ d , (3.10)where ∆ represents the spatial spac<strong>in</strong>g, as before, the values of j are def<strong>in</strong>ed byj ∈ [−L/∆, L/∆] d , with the spatial cutoff L chosen <strong>in</strong> such a way as to ensurethe localization of the evolution matrix with<strong>in</strong> the <strong>in</strong>terval [−L, L] d , and transitionamplitudes for small β can be calculated directly us<strong>in</strong>g the effective action approach[10, 44].65


3. Rotat<strong>in</strong>g ideal BEC3.2 F<strong>in</strong>ite number of energy eigenvalues and semiclassicalcorrectionsIn the previous section we have described a numerical approach that is capableof provid<strong>in</strong>g a large number of accurate energy eigenvalues for a general quantumsystem. For <strong>in</strong>stance, we are able to calculate typically 10 4 energy eigenvalues forthe considered BEC potential (3.4). In this section we discuss <strong>in</strong> more detail howthe f<strong>in</strong>iteness of numerically available energy eigenstates affects the calculation ofthermodynamic properties of Bose-E<strong>in</strong>ste<strong>in</strong> condensates.As outl<strong>in</strong>ed at the beg<strong>in</strong>n<strong>in</strong>g of this Chapter, the <strong>in</strong>formation on s<strong>in</strong>gle-particleeigenvalues is sufficient for calculat<strong>in</strong>g the condensation temperature accord<strong>in</strong>g toEq. (3.9). Below the condensation temperature, the ground-state occupancy followsfrom solv<strong>in</strong>g Eq. (3.8). In practical calculations, however, one is <strong>in</strong>evitably forcedto restrict the sum over j <strong>in</strong> the cumulant expansion (3.8) to some f<strong>in</strong>ite cutoff J,result<strong>in</strong>g <strong>in</strong> the follow<strong>in</strong>g approximation for the number of thermal atomsN − N 0 ≈J∑ ∞∑e −jβ(En−E0) . (3.11)j=1 n=1Thus, the ground-state occupancy N 0 depends not only on the particle number Nand the temperature T, but also on the cumulant cutoff J. In particular, whenwe solve Eq. (3.11) for the (<strong>in</strong>verse) condensation temperature β c , obta<strong>in</strong>ed fromthe condition N 0 = 0, we will get the solution <strong>in</strong> the form β c (J), with an explicitdependence on J. The exact condensation temperature β c is only obta<strong>in</strong>ed <strong>in</strong> thelimit J → ∞.Fig. 3.3 illustrates the J-dependence result<strong>in</strong>g from Eq. (3.11) for both a nonrotat<strong>in</strong>gand a critically rotat<strong>in</strong>g condensate. As expected, the sum saturates forhigh values of J to some f<strong>in</strong>ite number N − N 0 . By tun<strong>in</strong>g the temperature <strong>in</strong> sucha way that the sum saturates at the desired value of the total number of atoms N<strong>in</strong> the system, which corresponds to N 0 = 0, one is, <strong>in</strong> pr<strong>in</strong>ciple, able to extract thevalue of the condensation temperature T c .Although the results <strong>in</strong> Fig. 3.3 suggest that this approach can be appliedstraightforwardly, a closer look at the results for numerically calculated values ofN − N 0 reveals several problems that have to be addressed. At first we have to<strong>in</strong>vestigate how the results depend on the number of energy eigenstates used <strong>in</strong> the66


3. Rotat<strong>in</strong>g ideal BEC3.0·10 52.9·10 52.8·10 50 5 10 15 20 25 30N - N 02.7·10 52.6·10 52.5·10 52.4·10 5η=1, T=63.30 nKη=0, T=105.18 nKFigure 3.3: Number of thermally excited atoms N−N 0 as a function of the cutoff J <strong>in</strong>the cumulant expansion (3.11). The results are given for a non-rotat<strong>in</strong>g condensateat T = 105.18 nK and for a critically rotat<strong>in</strong>g condensate at T = 63.30 nK. Theresults are obta<strong>in</strong>ed by level p = 21 effective action, and all available numericaleigenstates are used to calculate N − N 0 . The discretization parameters were L =14.2 for η = 0 and L = 22.3 for η = 1. In both cases the spac<strong>in</strong>g was chosenaccord<strong>in</strong>g to L/∆ = 100, and the propagation time was t = 0.2.Jnumerical calculation. Fig. 3.4 gives this dependence for a critically rotat<strong>in</strong>g condensateat its critical temperature T c = 63.30 nK. We can see that the dependence onthe maximal available two-dimensional energy eigenvalue E max is quite significant.The <strong>in</strong>set of this figure reveals another problem: the value to which number N 0 −Nsaturates depends <strong>in</strong> addition on the cumulant cutoff J, as expla<strong>in</strong>ed earlier. Whilethe J-dependence can be dealt with by us<strong>in</strong>g a very large value of the cumulantcutoff <strong>in</strong> numerical calculations, the dependence on the maximal energy eigenvalueE max must be elim<strong>in</strong>ated by tak<strong>in</strong>g <strong>in</strong>to account a proper semiclassical correction tothe s<strong>in</strong>gle-particle partition functions.Namely, the f<strong>in</strong>ite number of energy eigenstates implies that the s<strong>in</strong>gle-particlepartition functions are only estimated byZ 1 (β) ≈n∑maxn=0e −βEn , (3.12)where n max corresponds to the value E max of the numerically available maximal en-67


3. Rotat<strong>in</strong>g ideal BECN - N 03.0·10 5 J=302.5·10 5J=10002.0·10 51.5·10 53.00·10 51.0·10 52.95·10 50.5·10 52.90·10 580 90 100 110 120 130 14000 20 40 60 80 100 120 140E max / − hω ⊥Figure 3.4: Number of thermally excited atoms N − N 0 calculated as a function ofthe maximal available two-dimensional energy eigenvalue E max at T = 63.30 nK.The results are given for two different values of the cumulant cutoff J for a criticallyrotat<strong>in</strong>g condensate, with the same parameters as <strong>in</strong> Fig. 3.3. The horizontal l<strong>in</strong>ecorresponds to the number of atoms N = 3 · 10 5 <strong>in</strong> the experiment [12].ergy eigenvalue. A semiclassical correction to this value, can be calculated accord<strong>in</strong>gto Ref. [65] as∫∆Z 1 (β, E max ) =d⃗r d⃗p(2π) 3 e−βH(⃗r,⃗p) Θ(H(⃗r, ⃗p) − E max ) , (3.13)where H(⃗r, ⃗p) as before represents the classical Hamiltonian of the system, while Θdenotes the Heaviside step-function.For the trap potential (3.4), <strong>in</strong> z-direction we have a pure harmonic potential,which can be treated exactly. Therefore, we focus only on the two-dimensionalproblem <strong>in</strong> the x − y plane. In this case, the semiclassical correction for the s<strong>in</strong>gleparticlepartition function (3.13) can be expressed <strong>in</strong> terms of the complementaryerror function:∆Z (2)1 (β, E max) = 1 { e−βE max2β κ√ π+κβ e β(1−η2 ) 24κ[−(1 − η 2 ) √ (1 − η 2 ) 2 + 4κE max]× Erfc(√βE max + β(1 − η2 ) 24κ)},(3.14)68


3. Rotat<strong>in</strong>g ideal BEC3.0·10 5 0 20 40 60 80 100 120 1402.5·10 5With SC corr.Without SC corr.2.0·10 5N - N 01.5·10 51.0·10 50.5·10 53.00·10 52.96·10 53.04·10 5 0 20 40 60 80 100 120 1400E max / − hω ⊥Figure 3.5: Number of thermally excited atoms N − N 0 calculated as a function ofE max with and without semiclassical corrections, calculated with a large cumulantcutoff J = 10 4 to elim<strong>in</strong>ate the J-dependence. The results correspond to a criticallyrotat<strong>in</strong>g condensate with the same parameters as <strong>in</strong> Fig. 3.3. The horizontal l<strong>in</strong>ecorresponds to N = 301834 which represents the exact value at T c = 63.30 nK.where the superscript denotes that only the x−y part of the potential is considered.When this semiclassical correction is taken <strong>in</strong>to account, the numerical resultsshow almost no dependence on E max , as can be seen from Fig. 3.5. Here we haveused an excessively large value of the cumulant cutoff J = 10 4 <strong>in</strong> order to completelyelim<strong>in</strong>ate any J dependence. From the <strong>in</strong>set <strong>in</strong> this graph we also see that E maxmust be chosen <strong>in</strong> accordance with the value estimated <strong>in</strong> the previous section forthe maximal reliable energy eigenvalue obta<strong>in</strong>ed by numerical diagonalization. Ifwe use a value E max larger than this, we will be underestimat<strong>in</strong>g the higher part ofthe energy spectra, and obta<strong>in</strong> <strong>in</strong>correct results. For a critically rotat<strong>in</strong>g condensatewith the anharmonicity κ = κ BEC the estimated value of E max from Table 3.3 isaround 90 ω, which agrees with the results from the <strong>in</strong>set of Fig. 3.5. If we use thisvalue for E max and calculate properties of the condensate us<strong>in</strong>g numerically obta<strong>in</strong>edeigenstates below E max with semiclassical corrections accord<strong>in</strong>g to Eq. (3.14), we willobta<strong>in</strong> the exact results with very high accuracy.69


3. Rotat<strong>in</strong>g ideal BEC3.3 Global properties of rotat<strong>in</strong>g BECsIn this section we will apply the presented approach to calculate different globalproperties of rotat<strong>in</strong>g BECs. First, we will calculate condensation temperature,and then, we will present phase diagrams def<strong>in</strong>ed <strong>in</strong> terms of condensate-fractiondependence on the temperature for different trap parameters. Additionally, wewill compare numerical results with semiclassical values and identify when the fullnumerical treatment becomes necessary.3.3.1 Condensation temperatureIf we take <strong>in</strong>to account semiclassical corrections, as expla<strong>in</strong>ed <strong>in</strong> the previous section,we can calculate, for <strong>in</strong>stance, the condensation temperature of the condensatefor different rotation frequencies. This implies that we have to f<strong>in</strong>d the temperaturefor which the number of thermal atoms saturates precisely at the total number ofatoms N. In practice, this works the other way around: for a given condensationtemperature T c we numerically calculate the particle number <strong>in</strong> the system us<strong>in</strong>gEq. (3.11), which gives the number of atoms <strong>in</strong> the system required for a condensationtemperature to be equal to T c . This procedure is implemented <strong>in</strong> Fig. 3.6 forseveral values of the rotation frequency Ω <strong>in</strong> units of η = Ω/ω. For example, forT c = 63.14 nK we see that the correspond<strong>in</strong>g number of particles is N = 3 · 10 5 ,which co<strong>in</strong>cides with the value for a critically rotat<strong>in</strong>g condensate <strong>in</strong> the experimentof Dalibard and collaborators [12].In pr<strong>in</strong>ciple, such a procedure is only applicable for low-accuracy calculations ofthe critical temperature, s<strong>in</strong>ce otherwise one has to use very large values of the cutoffJ which would practically slow-down numerical calculations. If one is <strong>in</strong>terested <strong>in</strong>more precise results, a suitable J-dependence must be properly taken <strong>in</strong>to account.In order to be able to efficiently extract the correct value of β c , we will derive ananalytical estimate of the asymptotic error ∆β c = β c − β c (J), which is <strong>in</strong>troducedby the presence of the cutoff J. Note that always ∆β c > 0, s<strong>in</strong>ce β c (J) < β c has tocompensate the miss<strong>in</strong>g terms <strong>in</strong> the sum (3.11).If we <strong>in</strong>sert β c = β c (J) + ∆β c <strong>in</strong>to Eq. (3.11), the error ∆β c can considered tobe small for sufficiently large value of the cutoff J. By compar<strong>in</strong>g Eq. (3.11) with70


3. Rotat<strong>in</strong>g ideal BEC4.0·10 53.5·10 53.0·10 52.5·10 520 40 60 80 100 120N - N 02.0·10 51.5·10 51.0·10 50.5·10 50η=1.0η=0.8η=0.6η=0.4η=0.0T [nK]Figure 3.6: Number of thermally excited atoms N −N 0 as a function of the temperatureT for different values of the rotation frequency and the quartic anharmonicityκ = κ BEC . The discretization parameters are given <strong>in</strong> Table 3.3, and the results arecalculated by tak<strong>in</strong>g <strong>in</strong>to account semiclassical corrections. The dashed l<strong>in</strong>e correspondsto the number of atoms N = 3 · 10 5 <strong>in</strong> the experiment [12]. For comparison,the full l<strong>in</strong>es depict the semiclassical results from Ref. [65].the exact expression (3.9) we obta<strong>in</strong>∞∑j=J+1 n=1∞∑e −jβc(En−E0) ≈ ∆β cJ∑j=1 n=1∞∑j(E n − E 0 ) e −jβc(En−E0) . (3.15)The term j(E n −E 0 ) with<strong>in</strong> the sum can be obta<strong>in</strong>ed by sett<strong>in</strong>g N 0 = 0 and apply<strong>in</strong>gthe partial derivative ∂/∂β c to Eq. (3.11):[ ]∂N ∂ ∑ ∞−∆β c ≈ 1 − ∆β c∂β c ∂β cj=J+1 n=1∞∑e −jβc(En−E0) . (3.16)Note that the derivative of the particle number N with respect to β c is not equal tozero, s<strong>in</strong>ce N is here effectively def<strong>in</strong>ed by the sum (3.9). Therefore, we have <strong>in</strong>stead∂N∂β c= −∞∑j=1 n=1∞∑j (E n − E 0 ) e −jβc(En−E0) . (3.17)Clearly, the right-hand side is a negative quantity that does not depend on J. How-71


3. Rotat<strong>in</strong>g ideal BECever, it does depend on β c , as well as on the energy spectrum of the system.If the system is close to a d-dimensional harmonic oscillator, which is the casefor the potential (3.4) with the small anharmonicity relevant for the experiment, forlarge values of J we have approximately∞∑∞∑j=J+1 n=1e −jβc(En−E0) ≈ d e−(J+1)βcω, (3.18)1 − e−βcω where ω denotes an effective harmonic frequency and d is the dimensionality ofthe correspond<strong>in</strong>g system. In our case, we apply the semi-classical correction onlyto the x − y part of the potential and for this reason d = 2. For the case of alarge anharmonicity, the effective frequency ω would depend on κ, represent<strong>in</strong>g theharmonic expansion of the potential around its m<strong>in</strong>imum. With such an estimate,Eq. (3.16) reduces tod∆β c ≈ −1 − e × e −(J+1)βcω. (3.19)−βcω ∂N/∂β c + (J + 1) e −(J+1)βcω dω1−e −βcωThe term (J+1) e −(J+1)βcω <strong>in</strong> the denom<strong>in</strong>ator of the second factor can be neglectedfor large enough values of the cutoff J, yield<strong>in</strong>g as a simplified version of the aboveexpression:d e −(J+1)βcω∆β c ≈ −∂N/∂β c (1 − e −βcω ) . (3.20)In order to use the derived estimates for ∆β c , apparently one would alreadyhave to know the sought-after value of β c as well as the difficult derivative ∂N/∂β c .However, <strong>in</strong> practical applications this obstacle can be circumvented as follows.The expressions (3.19) and (3.20) can be used for fitt<strong>in</strong>g the numerical data forβ c (J) = β c −∆β c , as is illustrated <strong>in</strong> Fig. 3.7. In this standard approach, all unknownvalues are fit parameters, obta<strong>in</strong>ed numerically by the least-square method. Notethat not only β c is obta<strong>in</strong>ed by such a fitt<strong>in</strong>g procedure, but also other parameters,such as ∂N/∂β c , or the effective harmonic frequency ω. The important po<strong>in</strong>t here isto capture the correct J-dependence, while all other parameters do not depend onit, so that they can be extracted by fitt<strong>in</strong>g. For example, <strong>in</strong> Fig. 3.7 we have used72


3. Rotat<strong>in</strong>g ideal BECβ c (J)0.054840.054820.054800.054780.054760.054740.054720.054700.054680.05466κ BEC , η=1.04, N=3·10 50.0546450 100 150 200 250 300JFigure 3.7: Dependence of β c on the cumulant cutoff J for an over-critically (η =1.04) rotat<strong>in</strong>g condensate of N = 3·10 5 atoms of 87 Rb with the quartic anharmonicityof the trap κ = κ BEC . The discretization parameters are given <strong>in</strong> Table 3.3. Thedashed l<strong>in</strong>e corresponds to a value of β c obta<strong>in</strong>ed by fitt<strong>in</strong>g the numerical results tothe function (3.21), while the full l<strong>in</strong>e gives the fitted function f(J).the fitt<strong>in</strong>g functionf(J) = β c −c 1 e −c 2(J+1)1 + c 3 (J + 1) e −c 4(J+1) , (3.21)which reproduces the numerical data quite accurately and gives high-precision resultsfor the condensation temperature T c . The virtue of the derived estimates lies<strong>in</strong> the fact that they can be used to extract the <strong>in</strong>formation on the condensationtemperature even for moderate values of J, when a saturation is not yet achieved.This substantially speeds up the numerical calculation of condensation temperatures,especially when it has to be done for different values of potential parameters,such as the frequency ratio η = Ω/ω.Fig. 3.8 summarizes the numerical results for the condensation temperature T cfor the anharmonicity κ = κ BEC as well as the particle numbers N = 3 · 10 5 andN = 1 · 10 4 . If we compare the obta<strong>in</strong>ed numerical results with the semiclassicalapproximation from Ref. [65], we see that the agreement turns out to be relativelygood for the undercritical regime, but it becomes worse for an overcritical rotation ofthe condensate. After present<strong>in</strong>g results for the ground-state occupancy, which were73


3. Rotat<strong>in</strong>g ideal BECT c [nK]12011010090807060504040353025201510κ BEC , N=1·10 4SC0 0.2 0.4 0.6 0.8 1ηκ BEC , N=3·10 50 0.2 0.4 0.6 0.8 1SCFigure 3.8: The condensation temperature as a function of the rotation frequencyfor the condensate of N = 3 · 10 5 and N = 1 · 10 4 atoms of 87 Rb, with the quarticanharmonicity of the trap κ = κ BEC . The discretization parameters are given <strong>in</strong>Table 3.3. The full l<strong>in</strong>es correspond to the semiclassical approximation for T c fromRef. [65].obta<strong>in</strong>ed from this approach <strong>in</strong> the next section, we will compare our numericallyexact results with the semiclassical approximation <strong>in</strong> more detail, and identify theparameter ranges where a full numerical treatment becomes necessary.3.3.2 Ground-state occupancyThe ground-state occupancy is the next important global property of Bose-E<strong>in</strong>ste<strong>in</strong>condensates we will look <strong>in</strong>to. Below the condensation temperature a non-trivialfraction of atoms is <strong>in</strong> the ground state, thus yield<strong>in</strong>g a macroscopic value of theoccupancy ratio N 0 /N.Us<strong>in</strong>g the same approach as above, we can calculate the ground-state occupancyfrom Eq. (3.8). After determ<strong>in</strong><strong>in</strong>g the ground-state energy E 0 from an exact diagonalizationof the evolution operator, we obta<strong>in</strong> the occupancy asN 0N = 1 − 1 N∞∑ [ejβE 0Z 1 (jβ) − 1 ] . (3.22)j=1In order to calculate N 0 /N, we need the full s<strong>in</strong>gle-particle energy spectrum. For74


3. Rotat<strong>in</strong>g ideal BEC10.8η=0, N=1·10 4η=1, N=5·10 4N 0 / N0.60.40.200 20 40 60 80 100 120 140T [nK]Figure 3.9: Ground-state occupancy N 0 /N as a function of the temperature T fornon-rotat<strong>in</strong>g and critically rotat<strong>in</strong>g condensate for different values of the total numberof atoms. The quartic anharmonicity is κ = 10 3 κ BEC , and the discretizationparameters are given <strong>in</strong> Table 3.3. The full l<strong>in</strong>es depict the semiclassical results fromRef. [65].low temperatures, the large number of energy eigenstates obta<strong>in</strong>ed with<strong>in</strong> the exactdiagonalization is sufficient. In the sum of Eq. (3.22) we have aga<strong>in</strong> to <strong>in</strong>troduce acutoff J and to elim<strong>in</strong>ate it by apply<strong>in</strong>g the methods described <strong>in</strong> previous sections.To this end one uses either a very large value for the cutoff or one derives theappropriate f<strong>in</strong>ite correction term, and fits the results to the derived function.Fig. 3.9 presents numerical results for the ground-state occupancy of the condensate.The quartic anharmonicity of the trap is chosen to be κ = 10 3 κ BEC , and theresults are given for the non-rotat<strong>in</strong>g case with the total number of atoms N = 1·10 4and for critically rotat<strong>in</strong>g condensate with N = 5 · 10 4 atoms. A comparison withthe semiclassical results derived <strong>in</strong> Ref. [65] shows that the deviations <strong>in</strong>crease withlarger temperatures and smaller particle numbers.3.3.3 Comparison with semiclassical approximationIn order to access the applicability and quality of the semiclassical approximation, wenow make a quantitative comparison of the semiclassical results and full numericalresults obta<strong>in</strong>ed <strong>in</strong> previous sections. Fig. 3.10 depicts the relative errors of thesemiclassically calculated condensation temperature. As we can see, the agreement is75


3. Rotat<strong>in</strong>g ideal BEC3.533210 3 κ BEC , N=3· 10 52.51(T SCc - T c ) / T c [%]21.5100 0.2 0.4 0.6 0.8 10.5κ BEC , N=3· 10 500 0.2 0.4 0.6 0.8 1η(T SCc - T c ) / T c [%]2015105765410 3 κ BEC , N=1· 10 40 0.2 0.4 0.6 0.8 10κ BEC , N=1· 10 40 0.2 0.4 0.6 0.8 1ηFigure 3.10: Relative error of semiclassical results for the condensation temperature[65] as a function of the rotation frequency Ω <strong>in</strong> units of η = Ω/ω for N = 3 · 10 5(top) and N = 1 · 10 4 (bottom). The quartic anharmonicity is κ = κ BEC , and thediscretization parameters are given <strong>in</strong> Table 3.3. The <strong>in</strong>sets <strong>in</strong> both plots give thecorrespond<strong>in</strong>g results for the large anharmonicity κ = 10 3 κ BEC .relatively good for large particle numbers and small anharmonicity if the condensaterotates under-critically. The error <strong>in</strong> this case is of the order of 1 % to 1.5 %, andturns out to be m<strong>in</strong>imal for a critical rotation. However, the error significantly<strong>in</strong>creases for an overcritical rotation up to almost 3.5 % for η = 1.1. Therefore,while the semiclassical approximation is acceptable for undercritical rotation, <strong>in</strong>the overcritical regime a numerical treatment becomes necessary. This is even more76


3. Rotat<strong>in</strong>g ideal BECpronounced if we decrease the particle number to 10 4 , which is quite typical for manyBEC experiments. In that case, semiclassical results already have an error of theorder of 20 %. For large anharmonicity the rotation effect is not so important, as wecan see from <strong>in</strong>sets on both graphs <strong>in</strong> Fig. 3.10. However, for the particle number 10 4a numerical treatment is <strong>in</strong>dispensable, s<strong>in</strong>ce the errors of the semiclassical resultsamount up to 5 %.3.4 Local properties of rotat<strong>in</strong>g BECsLocal properties of ultra-cold quantum gases are ubiquitously used to observe andstudy the phenomenon of Bose-E<strong>in</strong>ste<strong>in</strong> condensation. The prom<strong>in</strong>ent peak <strong>in</strong> TOFabsorption pictures, which appears suddenly when the temperature is decreasedbelow T c , is a clear signature for the occurrence of a BEC phase transition. It isexperimentally used to measure the thermodynamic properties of the condensate.In this section we will show how the presented numerical approach can be appliedto calculate both the density profiles and the TOF absorption imag<strong>in</strong>g profiles.3.4.1 Density profilesFor the ideal Bose gas, the density profiles of the condensate and of the gas phaseare given by Eqs. (1.11) and (1.12), respectively. Hav<strong>in</strong>g at our disposal numericallycalculated energy eigenvalues and eigenfunctions, we can calculate the density profileof the condensate. In order to do so, we first have to obta<strong>in</strong> the ground-stateoccupancy number N 0 us<strong>in</strong>g the approach described <strong>in</strong> the previous section. Oncethis is done, Eq. (1.12) allows to calculate the density profile. In view of a comparisonwith absorption imag<strong>in</strong>g, which always produces two-dimensional profiles, we have to<strong>in</strong>tegrate our numerically determ<strong>in</strong>ed three-dimensional particle density n(⃗r) alongthe imag<strong>in</strong>g axis. Fig. 3.11 presents typical results for the result<strong>in</strong>g density profilesof Bose-E<strong>in</strong>ste<strong>in</strong> condensates for both the non-rotat<strong>in</strong>g and the critically-rotat<strong>in</strong>gcase. Obviously, a rotation of the condensate leads to an effective spread<strong>in</strong>g due tothe appearance of a centrifugal potential.Although this approach is sufficient for treat<strong>in</strong>g the low-temperature regime,where the condensate is present, we emphasize that the same method can also beused to deal with the thermal regime, when the temperature is <strong>in</strong>creased above T c .For even higher temperatures, when the number of energy eigenstates, that need to77


3. Rotat<strong>in</strong>g ideal BECn(x, y)10·10 48·10 46·10 44·10 410·10 48·10 46·10 44·10 42·10 404·10 40-4-20x 24-4-202y4n(x, y)12·10 412·10 48·10 48·10 44·10 44·10 400-8-40x 48-8-404y8Figure 3.11: Density profile <strong>in</strong> xy-plane for a non-rotat<strong>in</strong>g (top) and a criticallyrotat<strong>in</strong>g (bottom) condensate of N = 3 · 10 5 atoms of 87 Rb with the anharmonicityκ = κ BEC at T = 30 nK. The dimensionless unit length on both graphs correspondsto 1.34 µm, i.e. the l<strong>in</strong>ear size of the profile is approximately 16.1 µm (top) and32.2 µm (bottom). The discretization parameters are given <strong>in</strong> Table 3.3.be taken <strong>in</strong>to account, exceeds the number of numerically accessible eigenstates, thepresented approach can be extended <strong>in</strong> a similar way as the partition function wascalculated previously as a sum of diagonal transition amplitudes. Us<strong>in</strong>g the cumulantexpansion of occupancies and the spectral decomposition of thermal transitionamplitudes, the density profile can be written for high enough temperatures asn(⃗r) = N 0 |ψ 0 (⃗r)| 2 + ∑ j≥1[ejβE 0A(⃗r,⃗r; jβ) − |ψ 0 (⃗r)| 2] . (3.23)Here A(⃗r,⃗r; jβ) represents the imag<strong>in</strong>ary-time amplitude for a s<strong>in</strong>gle-particle tran-78


3. Rotat<strong>in</strong>g ideal BECsition from the position ⃗r to the position ⃗r for the imag<strong>in</strong>ary time t = jβ.While both def<strong>in</strong>itions (1.12) and (3.23) are mathematically equivalent when oneis able to exactly calculate <strong>in</strong>f<strong>in</strong>itely many energy eigenstates and required transitionamplitudes for an arbitrary propagation time, the first def<strong>in</strong>ition is more suitable forlow temperatures, when the number of relevant energy eigenstates is moderate, andthe second one is suitable for high temperatures, when the imag<strong>in</strong>ary propagationtime β is small, and the short-time expansion can be successfully applied.3.4.2 Time-of-flight graphs for BECsIn typical BEC experiments, a trapp<strong>in</strong>g potential is switched off and the gas isallowed to expand freely dur<strong>in</strong>g a short flight time t which is of the order of severaltens of milliseconds. Afterwards an absorption picture is taken which maps thedensity profile to the plane perpendicular to the laser beam. For the ideal Bosecondensate, the density profile after time t is given byn(⃗r, t) = N 0 |ψ 0 (⃗r, t)| 2 + ∑ n≥1B n (E 0 , T)|ψ n (⃗r, t)| 2 , (3.24)where the density profile has to be <strong>in</strong>tegrated along the imag<strong>in</strong>g axis, and the eigenstatesψ n (⃗r, t) are propagated accord<strong>in</strong>g to the free Hamiltonian, conta<strong>in</strong><strong>in</strong>g only thek<strong>in</strong>etic term, s<strong>in</strong>ce the trapp<strong>in</strong>g potential is switched off. If the energy eigenstatesare available exactly, either analytically or numerically, their propagation <strong>in</strong> timecan be calculated by perform<strong>in</strong>g two consecutive Fourier transformations:ψ n (⃗r, t) =∫ d ⃗ k d ⃗ R(2π) 3 ei[⃗ k·(⃗r− ⃗ R)−ω ⃗k t] ψ n ( ⃗ R) , (3.25)where the term e −iω ⃗ k t accounts for a free-particle propagation <strong>in</strong> ⃗ k-space, ω ⃗k = 2 k 2 /(2M). In practical applications, when the energy eigenstates are calculatedby a numerical diagonalization of space-discretized transition amplitudes, the naturalway to calculate the above free-particle time evolution is to use Fast FourierTransform (FFT) numerical libraries.For high temperatures we can use a mathematically equivalent def<strong>in</strong>ition of thedensity profile which is derived aga<strong>in</strong> from us<strong>in</strong>g the cumulant expansion of occu-79


3. Rotat<strong>in</strong>g ideal BECpancy numbers and the spectral decomposition of transition amplitudes:n(⃗r, t) = N 0 |ψ 0 (⃗r, t)| 2 + ∑ j≥1[ ∫ d 3⃗ k1 d 3 ⃗ k2 d 3 X1 ⃗ d 3 X2 ⃗e jβE 0(2π) 6]×e i[(⃗ k 1 − ⃗ k 2 )·⃗r− ⃗ k 1· ⃗X 1 + ⃗ k 2· ⃗X 2 −(ω ⃗k1 −ω ⃗k2 )t] × A( X ⃗ 1 , X ⃗ 2 ; jβ) − |ψ 0 (⃗r, t)| 2 .(3.26)In both approaches it is first necessary to calculate the ground-state energy E 0and the eigenfunction ψ 0 (⃗r), as well as the ground-state occupancy N 0 . If we relyon Eqs. (3.24) and (3.25) to calculate TOF graphs, we have to calculate as manyeigenstates as possible by numerical diagonalization. Conversely, if it is possible touse directly Eq. (3.26), we can apply the effective action short-time expansion ofthermal transition amplitudes. In both cases FFT is ideally suited for calculat<strong>in</strong>gTOF graphs.3.4.3 Overcritical rotationThe case of critical and overcritical rotation η ≥ 1 is realized <strong>in</strong> the Paris experimentby <strong>in</strong>troduc<strong>in</strong>g the anharmonic part of the potential (3.4), so that the condensateis conf<strong>in</strong>ed even when the harmonic part of the trapp<strong>in</strong>g potential is completelycompensated or overcompensated by the rotation. The experimental realization ofthis delicate balance was difficult to achieve, but nevertheless when the condensatewas successfully conf<strong>in</strong>ed while rotat<strong>in</strong>g over-critically, the measurements of itsproperties can be done us<strong>in</strong>g the standard techniques, <strong>in</strong>clud<strong>in</strong>g absorption imag<strong>in</strong>g.With<strong>in</strong> the semiclassical approach one has to carefully consider this situation, s<strong>in</strong>cethe chemical potential is def<strong>in</strong>ed by the m<strong>in</strong>imum of the potential, Eq. (1.14), andnow cannot be simply set to zero anymore [65]. In our numerical approach, however,the implementation of the methods described <strong>in</strong> previous sections is straightforwardeven for overcritical rotation. First one calculates energy eigenvalues and eigenstatesus<strong>in</strong>g exact diagonalization, yield<strong>in</strong>g negative values for the first several eigenstates.Table 3.2 shows the result<strong>in</strong>g energy spectrum of an overcritically rotat<strong>in</strong>g condensate(η = 1.04) for the experimental value of the anharmonicity κ BEC , as well as forthe case of large anharmonicity 10 3 κ BEC .Condensation temperature and other global properties as well as local propertiesof overcritically rotat<strong>in</strong>g condensates can also be calculated as before. Fig. 3.12 gives80


3. Rotat<strong>in</strong>g ideal BEC500040003000n(x, y)TOF = 0 ms50002500500040003000n(x, y)TOF = 8 ms500025002000020000100010000-15-10-5x051015-10-15-5051510y0-15-10-5x051015-10-15-5051510y500040003000n(x, y)TOF = 16 ms50002500500040003000n(x, y)TOF = 24 ms500025002000020000100010000-15-10-5x051015-10-15-5051510y0-15-10-5x051015-10-15-5051510y500040003000n(x, y)TOF = 40 ms50002500500040003000n(x, y)TOF = 60 ms500025002000020000100010000-15-10-5x051015-10-15-5051510y0-15-10-5x051015-10-15-5051510yFigure 3.12: Time-of-flight absorption density profiles <strong>in</strong> xy-plane for an overcriticallyrotat<strong>in</strong>g (η = 1.04) condensate of N = 3 · 10 5 atoms of 87 Rb with theanharmonicity κ = κ BEC at T = 30 nK. The flight time, is given at each plot. Thedimensionless unit length on all graphs corresponds to 1.34 µm and the l<strong>in</strong>ear sizeof profiles is approximately 53.6 µm. The discretization parameters are given <strong>in</strong>Table 3.3.the TOF absorption imag<strong>in</strong>g sequence <strong>in</strong> the xy-plane for an overcritically (η = 1.04)rotat<strong>in</strong>g Bose-E<strong>in</strong>ste<strong>in</strong> condensate with the anharmonicity κ = κ BEC and the particlenumber N = 3 · 10 5 at T = 30 nK, exhibit<strong>in</strong>g an <strong>in</strong>terest<strong>in</strong>g behavior. The <strong>in</strong>itialdensity profile has a m<strong>in</strong>imum at the orig<strong>in</strong>, due to the shape of the anharmonicpotential. The free expansion of the condensate leads to an <strong>in</strong>crease <strong>in</strong> the particledensity at the orig<strong>in</strong>, and only afterwards the condensate density profile expandsmonotonically. Fig. 3.13 presents the time dependence of the particle density at81


3. Rotat<strong>in</strong>g ideal BEC6·10 4 0 0.02 0.04 0.06 0.08 0.112·10 3 0 0.02 0.04 0.06 0.08 0.1n(0, 0)5·10 44·10 43·10 42·10 41·10 4010·10 38·10 36·10 34·10 32·10 30η=0.80η=1.00η=1.04TOF [ms]Figure 3.13: Condensate density at the orig<strong>in</strong> of x − y plane as a function of thetime of flight (TOF) for the condensate of N = 3 · 10 5 atoms of 87 Rb at T = 30 nKfor several rotation frequencies Ω <strong>in</strong> units of η = Ω/ω. The quartic anharmonicityis κ = κ BEC and the discretization parameters are given <strong>in</strong> Table 3.3.the orig<strong>in</strong> for vary<strong>in</strong>g rotation frequencies, parameterized by the ratio η = Ω/ω.We read off that approach<strong>in</strong>g the critical rotation slows down the expansion ofthe condensate. For overcritical rotation this is even more pronounced, due to theappearance of the peak <strong>in</strong> the particle density for the expansion time t > 0. Thisleads to an expansion that is typically an order of magnitude slower for the rotationwith η > 1.3.5 Conclusions and outlookIn this Chapter the exact diagonalization of a time evolution operator is appliedto the study of ideal Bose gases <strong>in</strong> the anharmonic trap. Earlier derived higherordereffective actions are used for an efficient numerical calculation of both globaland local properties of fast-rotat<strong>in</strong>g Bose-E<strong>in</strong>ste<strong>in</strong> condensates. To this end we havecalculated large numbers of s<strong>in</strong>gle-particle eigenvalues and eigenstates and us<strong>in</strong>g this<strong>in</strong>formation, we have obta<strong>in</strong>ed the condensation temperature and the ground-stateoccupancy of the condensate, as well as density profiles and TOF absorption graphs.We have have focused on a critical and an overcritical rotation and have found asubstantial <strong>in</strong>crease <strong>in</strong> the time scale for the expansion of the condensate after the82


3. Rotat<strong>in</strong>g ideal BECtrapp<strong>in</strong>g potential is switched off compared to the undercritical case. Further studyshould <strong>in</strong>corporate the effects of weak <strong>in</strong>teractions on the Bose-E<strong>in</strong>ste<strong>in</strong> condensation<strong>in</strong> the external potential (3.4), either <strong>in</strong> the mean-field framework or by the fullnumerical treatment. F<strong>in</strong>ally, we note that approach presented here can also beused for numerical studies of properties of rotat<strong>in</strong>g ultra-cold Fermi gases [75].83


Chapter 4Mean-field description of an <strong>in</strong>teract<strong>in</strong>g BECIn the previous Chapter we have considered BEC of an ideal Bose gas, and,despite neglect<strong>in</strong>g the <strong>in</strong>teractions, we had to apply sophisticated numerical methodsfor characterization of the BEC phase transition and properties of the condensate.Now we extend the description of a Bose gas to <strong>in</strong>clude <strong>in</strong>teractions. For an exactmicroscopic description of an <strong>in</strong>teract<strong>in</strong>g Bose gas, we need to consider a full manybodyHamiltonian. For systems with a dom<strong>in</strong>ant two-body contact <strong>in</strong>teraction, theHamiltonian has the form (1.24):∫Ĥ =d⃗r(− ˆψ † (⃗r) 22M ∇2 ˆψ(⃗r) + V (⃗r) ˆψ† (⃗r) ˆψ(⃗r) + g )2 ˆψ † (⃗r) ˆψ † (⃗r) ˆψ(⃗r) ˆψ(⃗r) . (4.1)In the case of ultra-cold atomic gases, the bosons are alkali atoms, which are notelementary particles. However, due to very low temperatures, the atoms can beconsidered to be always <strong>in</strong> their ground states and their <strong>in</strong>ternal structure can beneglected. The exact treatment of the system described by the Hamiltonian (4.1)is possible only by numerical simulations based on Monte Carlo approach [76, 77].Such simulations are ubiquitously time-consum<strong>in</strong>g and significantly limit the scopeof problems that can be addressed. For this reason, a number of approximativemethods have been developed. Simplified approximative approaches are much easierfor the numerical implementation and sometimes even provide analytical <strong>in</strong>sights<strong>in</strong>to the studied problem. Furthermore, very often, <strong>in</strong>formation on the parametersof the system, such as the temperature or condensate fraction, are extracted fromexperimental data <strong>in</strong>directly, by fitt<strong>in</strong>g to expressions derived with<strong>in</strong> one of simplifiedschemes.In this Chapter we review the mean-field Hartree-Fock (HF) approximation fora many-body Hamiltonian (4.1). The HF approximation is the most-widely usedapproach to study f<strong>in</strong>ite-temperature properties of weakly <strong>in</strong>teract<strong>in</strong>g BECs [78, 5].We will show with<strong>in</strong> this mean-field scheme how the presence of the <strong>in</strong>teraction<strong>in</strong>fluences properties of the condensate, both <strong>in</strong> the zero-temperature limit, and <strong>in</strong>84


4. Mean-field description of an <strong>in</strong>teract<strong>in</strong>g BECthe vic<strong>in</strong>ity of the BEC phase transition. We focus on harmonically trapped bosonsfor two reasons: the trapp<strong>in</strong>g is always present <strong>in</strong> the experiments, and also thecalculation of the condensation temperature of the homogenous system turned out tobe a notorious problem, debated <strong>in</strong> many ways for several decades [79]. We scrut<strong>in</strong>yand compare several widely used implementations of the HF approximation with theemphasis on their advantages and drawbacks. The <strong>in</strong>terest <strong>in</strong> the topic <strong>in</strong>creases asnew experiments have reported the observation of beyond mean-field effects [80, 81],which should be <strong>in</strong>corporated <strong>in</strong>to the exist<strong>in</strong>g models.The partition function of the system <strong>in</strong> the grand canonical ensemble is given byZ(β) = Tr e −β(Ĥ−µ ˆN) , (4.2)and can be rewritten as a bosonic functional <strong>in</strong>tegral <strong>in</strong> the imag<strong>in</strong>ary time [32]:∮Z(β) =∮DΨDΨ ∗ e −A E[Ψ(⃗r,τ),Ψ ∗ (⃗r,τ)]/ , (4.3)where Ψ(⃗r, τ) and Ψ ∗ (⃗r, τ) are periodic functions, with the period β:Ψ(⃗r, τ) = Ψ(⃗r, τ + β) , Ψ ∗ (⃗r, τ) = Ψ ∗ (⃗r, τ + β) . (4.4)For the Hamiltonian (4.1) the Euclidean action A E is given byA E [Ψ(⃗r, τ), Ψ ∗ (⃗r, τ)] =∫ β0+ g 2∫dτ∫ β0(d⃗r Ψ ∗ (⃗r, τ) ∂)∂τ − 22M △ + V (⃗r) − µ Ψ(⃗r, τ)∫d⃗r Ψ ∗ (⃗r, τ)Ψ(⃗r, τ)Ψ ∗ (⃗r, τ)Ψ(⃗r, τ) . (4.5)dτA presence of the <strong>in</strong>teract<strong>in</strong>g Ψ 4 term <strong>in</strong> the action makes the calculation of the partitionfunction analytically <strong>in</strong>tractable, and to proceed further we apply the standardmean-field approach. In order to study Bose-E<strong>in</strong>ste<strong>in</strong> condensation, accord<strong>in</strong>g tothe Bogoliubov prescription (1.27), we first decompose the field Ψ <strong>in</strong>to the orderparameter ψ(⃗r, τ), which corresponds to the macroscopic condensate wave-function,and fluctuations δψ(⃗r, τ):Ψ(⃗r, τ) = ψ(⃗r, τ) + δψ(⃗r, τ) . (4.6)In the functional formalism this represents a change of variables, and the action now85


4. Mean-field description of an <strong>in</strong>teract<strong>in</strong>g BECconta<strong>in</strong>s terms up to the 4 th order <strong>in</strong> δψ that we should <strong>in</strong>tegrate over. Withoutfurther approximations, however, this problem is equivalent to the orig<strong>in</strong>al functional<strong>in</strong>tegral. Numerous approximation techniques were developed to treat terms of the3 rd and 4 th order <strong>in</strong> different approximative ways [19, 82]. In the Hartree-Fock-Bogoliubov approach, we approximate the 4 th order term asδψ ∗ δψ δψ ∗ δψ ≈4〈δψ ∗ δψ〉δψ ∗ δψ + 〈δψ ∗ δψ ∗ 〉δψδψ + 〈δψδψ〉δψ ∗ δψ ∗−2〈δψ ∗ δψ〉〈δψ ∗ δψ〉 − 〈δψδψ〉〈δψ ∗ δψ ∗ 〉. (4.7)In accordance with the previous decomposition, we <strong>in</strong>troduce auxiliary functionsh(⃗r, τ;⃗r, τ) = 〈δψ ∗ (⃗r, τ)δψ(⃗r, τ)〉 ,f(⃗r, τ;⃗r ′ , τ ′ ) = 〈δψ ∗ (⃗r, τ)δψ(⃗r ′ , τ ′ )〉 ,b(⃗r, τ;⃗r ′ , τ ′ ) = 〈δψ(⃗r, τ)δψ(⃗r ′ , τ ′ )〉 ,which are denoted as Hartree, Fock and Bogoliubov term, respectively. At themoment, the <strong>in</strong>troduced average values are purely formal, but can be later def<strong>in</strong>edso as to make the complete procedure self-consistent. In the case of the contact<strong>in</strong>teraction (1.21), the Hartree and Fock terms yield equal contributions, hence afactor of 4 <strong>in</strong> front of the correspond<strong>in</strong>g term <strong>in</strong> Eq. (4.7).After apply<strong>in</strong>g the mean-field approximation (4.7), the action A E becomes quadratic<strong>in</strong> δψ, and now the functional <strong>in</strong>tegrations of the Gaussian <strong>in</strong>tegrals can be explicitlyperformed. The f<strong>in</strong>al result for the partition function can be written <strong>in</strong> theformZ(β) = e −β Γ eff[ψ,ψ ∗ ,h,f,b] , (4.8)where Γ eff is the effective action, def<strong>in</strong>ed as a functional of five arguments: ψ(⃗r, τ),ψ ∗ (⃗r, τ), h(⃗r, τ;⃗r, τ), f(⃗r, τ;⃗r ′ , τ ′ ) and b(⃗r, τ;⃗r ′ , τ ′ ). They are determ<strong>in</strong>ed by extremiz<strong>in</strong>gthe effective action Γ eff [ψ, ψ ∗ , h, f, b] with respect to each of them:δΓ effδψ = 0, δΓ effδψ ∗= 0, δΓ effδh = 0, δΓ effδf = 0, δΓ effδb= 0.In the quest for the simplest mean-field description of an <strong>in</strong>homogeneous BEC,we will make another simplification by neglect<strong>in</strong>g Bogoliubov terms, i.e. anomalouscorrelations b(⃗r, τ;⃗r ′ , τ ′ ), as discussed <strong>in</strong> Ref. [83]. This assumption is justified <strong>in</strong>86


4. Mean-field description of an <strong>in</strong>teract<strong>in</strong>g BECthe limit T → 0, however it is very often used for all temperatures. The topic isexplored <strong>in</strong> detail <strong>in</strong> Ref. [84] where the consequences of the approximation are thoroughlydiscussed. Additionally, we neglect the possible depletion of the condensateat the zero temperature, i.e. the depletion that arises due to <strong>in</strong>teractions, which isa reasonable approximation <strong>in</strong> the case of a weakly <strong>in</strong>teract<strong>in</strong>g gas.F<strong>in</strong>ally, after implement<strong>in</strong>g all the described steps, we arrive at the f<strong>in</strong>ite-temperatureHF description of a bosonic gas, which is given by the follow<strong>in</strong>g system ofequations:[ ∂]∂τ − 22M △ + V (⃗r) + g|ψ(⃗r, τ)|2 + 2 gh(⃗r, τ;⃗r, τ) ψ(⃗r, τ) = µψ(⃗r, τ) , (4.9)h(⃗r, τ;⃗r, τ) = ∑ ψ ⃗k (⃗r)ψ ∗ 1⃗ k(⃗r)e β(E ⃗ −µ) k − 1 , (4.10)⃗ k][− 22M △ + V (⃗r) + 2 g|ψ(⃗r, τ)|2 + 2 gh(⃗r, τ;⃗r, τ) ψ ⃗k (⃗r) = E ⃗k ψ ⃗k (⃗r) , (4.11)where ψ ⃗k (⃗r) are effective s<strong>in</strong>gle-particle wave-functions, and E ⃗k are the correspond<strong>in</strong>geigenvalues. More details on the derivation can be found <strong>in</strong> Ref. [78]. Althoughformally the HF equations depend on the imag<strong>in</strong>ary time τ, physically is only relevantthe equilibrium case, when the macroscopic wave-function of the condensateψ(⃗r, τ) does not depend on τ anymore (∂ψ(⃗r, τ)/∂τ = 0), but only on the position⃗r. The above set of equations has to be solved self-consistently, tak<strong>in</strong>g <strong>in</strong>to accountthat the total number of particles is fixed to N, and leads to the solution that consistsof the effective s<strong>in</strong>gle-particle eigenfunctions ψ ⃗k and eigenvalues E ⃗k , the Hartreefunction h, and the condensate wave-function ψ. From Eq. (4.10) we immediatelysee the physical <strong>in</strong>terpretation of the Hartree function h, which represents the densityof the thermal cloud, n th (⃗r). Effectively, with<strong>in</strong> the mean-field description, thegas of bosons is split <strong>in</strong>to the condensate and thermal component. We note the closeanalogy with the non<strong>in</strong>teract<strong>in</strong>g gas description presented <strong>in</strong> Chapter 1, with theimportant exception that the two components now mutually <strong>in</strong>teract. By vary<strong>in</strong>gthe total number of particles N and the temperature T, a complete N −T phase diagramcan be explored. Before consider<strong>in</strong>g a BEC phase transition <strong>in</strong> the mean-fieldapproximation, we first present the zero-temperature limit of the HF approximation.87


4. Mean-field description of an <strong>in</strong>teract<strong>in</strong>g BEC4.1 Gross-Pitaevskii equationIn the zero-temperature limit, we can neglect the thermal cloud, and setlim n th(⃗r) = 0 , (4.12)β→∞which stems directly from Eq. (4.10). In this case, Eq. (4.11) becomes irelevant,while from Eq. (4.9) we f<strong>in</strong>d the time-<strong>in</strong>dependent Gross-Pitaevskii (GP) equation[13, 14, 34] for the order parameter:[− 22M △ + V (⃗r) + g|ψ(⃗r)|2 ]ψ(⃗r) = µψ(⃗r) . (4.13)Effectively, <strong>in</strong> the mean-field approximation at zero temperature, we assume thatall atoms occupy the same state ψ(⃗r), which we denote as the condensate wavefunction.Note that already by neglect<strong>in</strong>g anomalous averages we have disregardeda depletion of the condensate at zero temperature that arises due to <strong>in</strong>teractions.However, it turns out that this is a reasonable approximation <strong>in</strong> the wide range ofexperimental parameters for a weakly <strong>in</strong>teract<strong>in</strong>g gas. On the left-hand side of theGP equation (4.13) we have a k<strong>in</strong>etic energy term, an external trap potential V (⃗r),and a nonl<strong>in</strong>ear term orig<strong>in</strong>at<strong>in</strong>g from the mean-field <strong>in</strong>teraction between the atoms.The GP equation belongs to the class of nonl<strong>in</strong>ear Schröd<strong>in</strong>ger equations, which areextensively studied also <strong>in</strong> the field of nonl<strong>in</strong>ear optics [85, 86].Now, let us analyze solutions of the GP equation. To beg<strong>in</strong> with, we note thatthe non<strong>in</strong>teract<strong>in</strong>g limit is straightforwardly reproduced: for g → 0, the condensatewavefunction is the ground state of the external potential V (⃗r), and the value of thechemical potential is equal to the ground-state energy. In the limit of strong repulsive<strong>in</strong>teractions (for a large number of atoms, for example), the term correspond<strong>in</strong>gto the k<strong>in</strong>etic energy can be safely neglected. In that case we f<strong>in</strong>d an algebraicstationary solution|ψ(⃗r)| 2 = 1 (µ − V (⃗r)) θ(µ − V (⃗r)), (4.14)gwhich is the well-known Thomas-Fermi (TF) solution [19]. The value of the chemicalpotential µ is determ<strong>in</strong>ed as usual, by fix<strong>in</strong>g the total number of atoms <strong>in</strong> the systemto N. In particular, for the harmonic trap, the solution for the non<strong>in</strong>teract<strong>in</strong>g case isa Gaussian, while the TF solution yields a parabolic profile. It is easy to understand88


4. Mean-field description of an <strong>in</strong>teract<strong>in</strong>g BEC0.450.40.35g = 0g = 125.484g = 627.40.3ψ(r/l)0.250.20.150.10.0500 1 2 3 4 5 6r/lFigure 4.1: Wave functions of the condensate for different <strong>in</strong>teraction strengths. Weassume a spherically symmetric harmonic trap V (⃗r) = Mω 2 r 2 /2. Radial coord<strong>in</strong>ateis expressed <strong>in</strong> units of l = √ /Mω, while the <strong>in</strong>teraction strength is given <strong>in</strong> thedimensionless units g ≡ 4πaN/l.that the repulsive <strong>in</strong>teraction leads to the broaden<strong>in</strong>g of the density profile comparedto the non<strong>in</strong>teract<strong>in</strong>g Gaussian. This is illustrated <strong>in</strong> Fig. 4.1, where we show wavefunctions of the condensate obta<strong>in</strong>ed by numerically solv<strong>in</strong>g the GP equation fordifferent <strong>in</strong>teraction strengths. To f<strong>in</strong>d the condensate ground state, we perform theimag<strong>in</strong>ary-time propagation of the GP equation [87]. More details on the numericalalgorithms are given <strong>in</strong> Appendix A. Due to its simplicity, TF approximation iswidely used for the <strong>in</strong>terpretation of experimental data obta<strong>in</strong>ed for systems with alarge number of atoms.In a similar manner, start<strong>in</strong>g from the real-time formalism and neglect<strong>in</strong>g thecondensate depletion, <strong>in</strong> the mean-field framework we obta<strong>in</strong> the time-dependentGP equation:]∂ψ(⃗r, t)i =[− 2∂t 2M ∆ + V (⃗r) + g|ψ(⃗r, t))|2 ψ(⃗r, t), (4.15)which describes the condensate dynamics at T = 0. This equation also allows thestudy of the condensate excitation spectra, which is essential <strong>in</strong>formation for prob<strong>in</strong>gsystem’s properties.The time-dependent GP equation can be also derived variationally, from the89


4. Mean-field description of an <strong>in</strong>teract<strong>in</strong>g BECaction pr<strong>in</strong>ciple [19] based on the Lagrangian∫L GP = d⃗r L GP (⃗r) , (4.16)where the Lagrangian density is given byL GP (⃗r) = i 2) (ψ ∂ψ∗∂t − ψ∗∂ψ + 2∂t 2M |∇ψ|2 + V (⃗r)|ψ| 2 + g 2 |ψ|4 . (4.17)By perform<strong>in</strong>g the extremization of the action with respect to ψ(⃗r, t),δ ∫ t 2t 1dtL GPδψ ∗ (⃗r, t)= 0, (4.18)we aga<strong>in</strong> arrive at the GP Eq. (4.15). This is an important observation, s<strong>in</strong>ce itrepresents the basis of the time-dependent variational analysis. We exploit thisapproach later <strong>in</strong> Chapter 5.4.2 F<strong>in</strong>ite-temperature properties of a BECAfter briefly explor<strong>in</strong>g the properties of a weakly <strong>in</strong>teract<strong>in</strong>g BEC <strong>in</strong> the zerotemperaturelimit, we now cont<strong>in</strong>ue the study of its f<strong>in</strong>ite-temperature properties<strong>in</strong> the HF approximation, Eqs. (4.9)-(4.11). To facilitate an almost analytical approach,we abolish Eq. (4.11) and treat excited states with<strong>in</strong> a semiclassical approximation,while keep<strong>in</strong>g the generalized GP equation (4.9) for the condensate phase.To apply the semiclassical approximation, similarly to derivation of Eq. (1.13), weuse the classical Hamiltonian H(⃗r, ⃗p) = ⃗p 2 /2M + 2g(n 0 (⃗r) + n th (⃗r)), <strong>in</strong> accordancewith Eq. (4.11), and replace a summation over the discrete eigenstates <strong>in</strong> Eq. (4.10)by an <strong>in</strong>tegration over momentum ⃗p. We assume a spherically symmetric trap,i.e. V (⃗r) = V (r), and search for the equilibrium configuration (∂ψ(⃗r, τ)/∂τ = 0),when ψ(⃗r, τ) ≡ ψ(r). Eqs. (4.10) and (4.11) are now reduced to a s<strong>in</strong>gle equation,and, together with the equation for a total number of particles, we have the follow<strong>in</strong>g90


system of equations:4. Mean-field description of an <strong>in</strong>teract<strong>in</strong>g BEC][− 22M △ + V (r) + g n 0(r) + 2 g n th (r) ψ(r) = µ ψ(r) , (4.19)n th (r) = 1 (ζ ) λ 3 3/2 eβ(µ−2g(n th (r)+n 0 (r))−V (r)),T∫ ∫(4.20)N = N 0 + N th = d⃗r n 0 (r) + d⃗r n th (r) . (4.21)Here, as before, the condensate density is given by n 0 (r) = |ψ(r)| 2 . We denote theabove system of equations as GPSC model.Before consider<strong>in</strong>g the full GPSC model of a BEC, we will first consider a seriesof simpler models: the almost-ideal model (widely used <strong>in</strong> the analysis of the experimentaldata), the semi-ideal model [88, 89], and the Thomas-Fermi-semiclassicalmodel (TFSC), <strong>in</strong> which Eq. (4.19) is solved <strong>in</strong> the TF approximation. These modelsgradually add more details to the description of a BEC and become more complex,while at the same time they allow us to study properties of a BEC <strong>in</strong> a systematicway and to build new knowledge based on the previously obta<strong>in</strong>ed one.In the rest of this Chapter, we assume an isotropic harmonic trap V (r) =Mω 2 r 2 /2, with a typical length scale l = √ /Mω. For 87 Rb we use the valueof the scatter<strong>in</strong>g length of a = 5.4 nm <strong>in</strong> all presented numerical results.4.2.1 Almost-ideal modelThe crudest f<strong>in</strong>ite-temperature approximation of an <strong>in</strong>teract<strong>in</strong>g BEC takes <strong>in</strong>to accountonly the effect of <strong>in</strong>teractions on the condensate cloud, while it considersthermal atoms as an ideal gas. This is justified by the fact that, due to a spatialcompression of the condensate <strong>in</strong> the external trap, the density of this componentis high and effects of <strong>in</strong>teraction play a prom<strong>in</strong>ent role. On the other hand, thermalcloud is more spread, with a lower density, which allows us to consider it as anideal gas. To simplify the model further, we describe the condensate with<strong>in</strong> the TFapproximation. Thus, us<strong>in</strong>g Eqs. (4.14) and (1.13), we obta<strong>in</strong>:n 0 (r) = 1 (µ − V (r))θ(µ − V (r)) ,g(4.22)n th (r) = 1 (ζ ) λ 3 3/2 eβ(µ−V (r)).T(4.23)91


4. Mean-field description of an <strong>in</strong>teract<strong>in</strong>g BECn 0 (r)n th (r)n(r)4⋅10 19 0 10 20 30 40 503⋅10 192⋅10 191⋅10 190r[µm]Figure 4.2: Density profiles of different components (m −3 ) versus the radial coord<strong>in</strong>ater with<strong>in</strong> the almost-ideal model for the parameters N = 10 6 , T = 60 nK,ω = 100 Hz.Obviously, <strong>in</strong> order to have a well-behaved value of n th (0), we have to use theapproximation µ = 0 for the thermal gas <strong>in</strong> the condensate phase. A typical densityprofiles are shown <strong>in</strong> Fig. 4.2. With<strong>in</strong> this framework, the two components do not<strong>in</strong>teract mutually and their density profiles are monotonously decreas<strong>in</strong>g functions ofthe radial coord<strong>in</strong>ate. Although the model is oversimplified and the thermodynamicproperties of a BEC it provides correspond to the non<strong>in</strong>teract<strong>in</strong>g gas, it has beenwidely used s<strong>in</strong>ce the first observation of a BEC until the present state-of-the-artexperiments.4.2.2 Semi-ideal modelA basic premise of this model is to treat thermal atoms as an ideal gas with<strong>in</strong> theeffective potential, which is a comb<strong>in</strong>ation of the external trap potential and themean-field repulsive <strong>in</strong>teraction of the condensate: V eff (r) = V (r)+2gn 0 (r) [88, 89].Additionally, we neglect the <strong>in</strong>fluence of the thermal component on the condensate,and describe the BEC ground-state with<strong>in</strong> the TF approximation. With this setof simplifications, from Eqs. (1.13) and (4.14), we obta<strong>in</strong> the follow<strong>in</strong>g system of92


4. Mean-field description of an <strong>in</strong>teract<strong>in</strong>g BECequations:n 0 (r) = 1 (µ − V (r))θ(µ − V (r)) ,g(4.24)n th (r) = 1 (ζ ) λ 3 3/2 e−β|µ−V (r)|.T(4.25)In the condensed phase, the value of the chemical potential is given by µ = ω 2 (15N 0al) 2 5,which is obta<strong>in</strong>ed by fix<strong>in</strong>g the number of atoms <strong>in</strong> BEC to N 0 . Typical densityprofiles obta<strong>in</strong>ed by solv<strong>in</strong>g Eqs. (4.24) and (4.25) are shown <strong>in</strong> Fig. 4.3. We observea spatial separation of the condensate and the thermal cloud - condensate resides<strong>in</strong> the trap center, while thermal atoms are more spread. Additionally, due to arepulsive effect of the condensate to the thermal component, the density profile ofthe thermal cloud is not a monotonous function of the radial coord<strong>in</strong>ate r, unlike <strong>in</strong>the case of the almost-ideal model. This is far more realistic from the experimentalpo<strong>in</strong>t of view.n 0 (r)n th (r)n(r)3⋅10 19 0 10 20 30 40 502⋅10 191⋅10 190r[µm]Figure 4.3: Density profiles of different components (m −3 ) versus the radial coord<strong>in</strong>ater with<strong>in</strong> the semi-ideal model for the parameters N = 10 6 , T = 60 nK,ω = 100 Hz.The semi-ideal model def<strong>in</strong>ed by Eqs. (4.24) and (4.25) is simple enough to evenallow an analytic calculation of thermodynamic properties of a BEC. It was shownthat it yields a lower condensate fraction with respect to the non<strong>in</strong>teract<strong>in</strong>g case(1.10) for the same temperature [88, 89]. However, it turns out that the condensationtemperature with<strong>in</strong> the semi-ideal approach is the same as for the ideal gas model.93


4. Mean-field description of an <strong>in</strong>teract<strong>in</strong>g BECn 0 (r)n th (r)3⋅10 19 0 10 20 30 40 502⋅10 191⋅10 190r[µm]Figure 4.4: Density profiles of different components (m −3 ) versus the radial coord<strong>in</strong>ater with<strong>in</strong> the TFSC model for the parameters N = 874403, T = 60 nK,ω = 100 Hz. Red solid curves are analytic solutions obta<strong>in</strong>ed us<strong>in</strong>g the Rob<strong>in</strong>sonformula.This is too crude approximation from the experimental po<strong>in</strong>t of view and, <strong>in</strong> orderto understand how the properties of a BEC phase transition are modified <strong>in</strong> thepresence of <strong>in</strong>teractions, we need to develop a better approximation.4.2.3 TFSC modelIn this model, we describe the ground state us<strong>in</strong>g the TF approximation, whileexcited states are treated with<strong>in</strong> the semiclassical approximation (4.20), withoutfurther simplifications. Hence, we arrive at the follow<strong>in</strong>g system of equations:n 0 (r) = 1 g (µ − V (r) − 2gn th(r)), (4.26)n th (r) = 1 (ζ ) λ 3 3/2 e−β|µ−V (r)−2gn th (r)|, (4.27)Twhich we call the Thomas-Fermi-semiclassical model (TFSC). A typical numericalsolution of Eqs. (4.26) and (4.27) is given <strong>in</strong> Fig. 4.4. The obta<strong>in</strong>ed density profilesare similar to those of the semi-ideal model. However, <strong>in</strong> this model certa<strong>in</strong> unphysicaldiscont<strong>in</strong>uities appear close to the condensate boundary, where we observejumps both <strong>in</strong> the density of the thermal cloud and of the condensate component.In order to verify the obta<strong>in</strong>ed numerical results, we will use an analytic approx-94


4. Mean-field description of an <strong>in</strong>teract<strong>in</strong>g BECimation of the TFSC model. To this end, we rewrite the semiclassical expression(4.27) us<strong>in</strong>g the Rob<strong>in</strong>son formula [63]:ζ ν (e z ) = Γ(1 − ν)(−z) ν−1 +∞∑k=0z kk! ζ ν−k, (4.28)where, as before, Γ(z) is the Gamma function, and ζ ν is the Riemann zeta function.Close to the condensate boundary, the value of µ − V (r) − 2gn th (r) is small and wecan rely on the approximationζ 3/2 (e z ) ≈ Γ(−1/2)(−z) 1/2 + ζ 3/2 + zζ 1/2 , (4.29)to obta<strong>in</strong> an implicit equation for n th (r):n th (r) ≈ 1 [Γ(−1/2) β 1/2 |µ − V (r) − 2gnλ 3 th (r)| 1/2T]+ζ 3/2 − ζ 1/2 β |µ − V (r) − 2gn th (r)| . (4.30)When solv<strong>in</strong>g Eq. (4.30), we encounter four different possible branches for n th (r),due to the two quadratic equations. However, only two branches are physicallymean<strong>in</strong>gful and can be identified easily. Results for the density of the thermalcomponent obta<strong>in</strong>ed <strong>in</strong> this way are also shown <strong>in</strong> Fig. 4.4 by red solid l<strong>in</strong>es. Wesee good agreement of the approximation (4.30) with numerically calculated valuesbased on the full TFSC model. Furthermore, this analytical approach confirms thepresence of discont<strong>in</strong>uities <strong>in</strong> the density profiles close to the condensate boundary.Therefore, the discont<strong>in</strong>uities are not an artifact of numerical simulations, but theproblem of the model itself. This issue was noticed already <strong>in</strong> the early Ref. [33],and it is a consequence of the TF approximation. However, the model is able topredict an approximate phase diagram of a BEC, which is <strong>in</strong> a very good agreementwith experimental data, as we discuss later <strong>in</strong> this Chapter.4.2.4 GPSC modelF<strong>in</strong>ally, we consider the full set of Eqs. (4.19)-(4.21), i.e. the GPSC model. We solvethe equations for a fixed number of particles N us<strong>in</strong>g an iterative method. Oneway to start the iterations is to assume that all atoms are <strong>in</strong> the condensate and tosolve Eq. (4.19) without the thermal cloud. In the next step, the previous solution95


4. Mean-field description of an <strong>in</strong>teract<strong>in</strong>g BEC2⋅10 19 0 10 20 30 40 50 60 70 801⋅10 19 26 27 28 29 30 31 321.5⋅10 198⋅10 18n th (r)1⋅10 195⋅10 180semi-idealTFSCGPSCr[µm]Figure 4.5: The density profile of thermal atoms n th (m −3 ) versus the radial coord<strong>in</strong>ater, calculated us<strong>in</strong>g the three different approximations for the values ofparameters N = 10 7 , T = 100 nK, ω = 100 Hz. We see a good agreement betweenthe TFSC and GPSC calculations, due to a large numbers of atoms <strong>in</strong> the trap.for the condensate cloud is <strong>in</strong>serted <strong>in</strong>to the algebraic Eq. (4.20), which is thensolved for n th (r). After <strong>in</strong>tegration of n th (r) over r, a new value of the number ofatoms <strong>in</strong> the condensate N 0 is calculated as N −N th , and the procedure is repeatedby solv<strong>in</strong>g Eq. (4.19), but with the new values of N 0 and n th (r), calculated <strong>in</strong> theprevious step. This procedure is repeated until the desired convergence of the resultsis achieved. Of course, many modifications of the described procedure are possible,and practically one would like to identify the iteration procedure which leads to thefastest convergence toward the f<strong>in</strong>al result. The obta<strong>in</strong>ed density profiles are shown<strong>in</strong> Fig. 4.5. We stress that the discont<strong>in</strong>uities present <strong>in</strong> the TFSC model are absent<strong>in</strong> the GPSC model. Obviously, the presence of the k<strong>in</strong>etic energy term <strong>in</strong> Eq. (4.19)leads to smooth density distributions and cures the discont<strong>in</strong>uities problem of theTFSC model. For a large number of particles, TFSC and GPSC solutions are veryclose to each other, while for the smaller number of particles the TFSC solution iscloser to the semi-ideal model result.To study onset of Bose-E<strong>in</strong>ste<strong>in</strong> condensation, let us consider the follow<strong>in</strong>g scenario:we fix the temperature T, and then change the total number of particles N <strong>in</strong>the trap. As the number of particles <strong>in</strong>creases, the occupation of the ground statebecomes higher and higher, and, above a certa<strong>in</strong> threshold, condensation sets <strong>in</strong>. Illustrativenumerical results are given <strong>in</strong> Fig. 4.6. A normalized density 4πn(r)r 2 /N96


4. Mean-field description of an <strong>in</strong>teract<strong>in</strong>g BEC4 π n(r) r 2 /N [m -1 ]35000300002500020000150001000050000N = 4 × 10 7N = 3 × 10 7N = 2 × 10 7N = 1.46 × 10 7N = 1.09 × 10 70 50 100 150 200r [µm]ω = 125.664 HzT = 200 nKFigure 4.6: Plot of 4πn(r)r 2 /N versus the radial coord<strong>in</strong>ate r with<strong>in</strong> the GPSCmodel for the parameters T = 200 nK, ω = 125.664 Hz.versus the radial coord<strong>in</strong>ate r with<strong>in</strong> the GPSC model is shown <strong>in</strong> the graph. Fora very large number of atoms we see a s<strong>in</strong>gle dom<strong>in</strong>ant peak correspond<strong>in</strong>g to thecondensate. As the number of atoms decreases, the thermal component becomesvisible, and another broader peak emerges. F<strong>in</strong>ally, below the critical value of thetotal number of atoms, the condensate component disappears completely.High accuracy of the GPSC model is confirmed by direct comparison with thenumerical results obta<strong>in</strong>ed from Monte Carlo simulations [76]. However, GPSCmodel fails precisely <strong>in</strong> describ<strong>in</strong>g the phase transition properly. It turns out thatthe system of Eqs. (4.19)-(4.21) does not admit a solution <strong>in</strong> the vic<strong>in</strong>ity of thephase transition, when it is approached from the condensate side. For all physicalquantities to be well-behaved, it is crucial that the value of the chemical potentialis lower than the value of the classical energy of excited states, µ < H(⃗r, ⃗p). Unfortunately,for small values of n 0 (r), this condition is not satisfied. The <strong>in</strong>consistencyarises s<strong>in</strong>ce we treat the ground state quantum-mechanically and all the other statessemiclassically. For this reason, <strong>in</strong> the next subsection, we calculate thermodynamicproperties of <strong>in</strong>teract<strong>in</strong>g bosons with<strong>in</strong> the TFSC model.4.2.5 Calculation of the condensation temperatureA representative phase diagram of ultra-cold gas of bosons <strong>in</strong> the harmonic trapis displayed <strong>in</strong> Fig. 4.7. Calculations are performed with<strong>in</strong> the TFSC model. We97


4. Mean-field description of an <strong>in</strong>teract<strong>in</strong>g BECN 0 /N10.80.60.4N=10 6 , idealN=10 7 , idealN=10 6 , semi-idealN=10 7 , semi-idealN=10 6 , TFSCN=10 7 , TFSC0.2020 40 60 80 100 120 140 160 180 200T [nK]Figure 4.7: Condensate fraction N 0 /N versus the temperature T for ω = 100 Hz.0.50.4N=10 7 , idealN=10 7 , TFSCN=10 7 , semi-idealN 0 /N0.30.20.10120 125 130 135 140 145 150 155 160T [nK]Figure 4.8: Condensate fraction N 0 /N versus the temperature T for ω = 100 Hz.In the semi-ideal model the condensation temperature is the same as <strong>in</strong> the idealcase T 0 c = 154.77 nK, while the TFSC approximation yields a lower condensationtemperature T c = 149.5 nK, <strong>in</strong> agreement with the analytical calculation.have solved Eqs. (4.26) and (4.27) for the fixed number of particles <strong>in</strong> the giventemperature range and calculated the correspond<strong>in</strong>g condensate fractions N 0 /N.From Fig. 4.8, we clearly see the decrease <strong>in</strong> the condensation temperature due to<strong>in</strong>teraction effects.We now focus on the analytical calculation of the shift <strong>in</strong> the condensation tem-98


4. Mean-field description of an <strong>in</strong>teract<strong>in</strong>g BECperature with respect to the non<strong>in</strong>teract<strong>in</strong>g case [90]. To study effects of <strong>in</strong>teractionwith<strong>in</strong> the TFSC model, we start from the gas phase. In that case the condensateis absent, and n 0 (r) = 0. If we <strong>in</strong>sert this <strong>in</strong>to Eq. (4.27), we getn th (r) = 1 (ζ ) λ 3 3/2 eβ(µ−2gn th (r)−V (r)). (4.31)TThe condensation first appears <strong>in</strong> the center of the trap, where the follow<strong>in</strong>g relationholds:n th (0) = 1 (ζ ) λ 3 3/2 eβ(µ−2gn th (0)).TTherefore, at the onset of a BEC phase transition, we have µ c = 2gn c th (0) andn c th (0) = 1 ζλ 3 3/2 , which yields the critical value of the chemical potential is µ c =T2gζ 3/2 /λ 3 T . Now we calculate the critical number of atoms <strong>in</strong> the trap as∫ ∞N c = 4π dr r 2 1 (ζ )0 λ 3 3/2 eβ(µ c−2gn th (r)−V (r)). (4.32)TTo calculate the above <strong>in</strong>tegral, we use the perturbative approach and expand thezeta function to power series <strong>in</strong> the <strong>in</strong>teraction constant g to l<strong>in</strong>ear terms. Thismeans that we consider the term µ c − 2gn th (r) as a small quantity. In this way weobta<strong>in</strong>N c ≈ 4π≈λ 3 T4πλ 3 T×= 1λ 3 T∫ ∞0∫ ∞0(2 gλ 3 Tdr r 2 [ ζ 3/2(e−βV (r) ) + βζ 1/2(e−βV (r) ) (µ c − 2gn th (r)) ]dr r 2 [ ζ 3/2(e−βV (r) ) + βζ 1/2(e−βV (r) )ζ 3/2 − 2 g (ζ ))]λ 3 3/2 e−βV (r)T[ ] 3/2 [ 2πζβMω 2 3 + 2gβ 1 ζλ 3 3/2 ζ 2 − 2gβ 1 ζTλ 3 T( 12 , 3 2 ; 3 2)], (4.33)where ζ(i, j; k) = ∑ ∞m,n=1 1/mi n j (m + n) k . From the above expression we immediatelyf<strong>in</strong>d the shift of the condensation temperature [90]:δT c = T c<strong>in</strong>t− T 0 cT 0 c= −1.33 a l N1/6 , (4.34)and conclude that repulsive <strong>in</strong>teractions lead to the lower condensation temperaturecompared to the non<strong>in</strong>teract<strong>in</strong>g case. The result (4.34) is derived by assum<strong>in</strong>g that99


4. Mean-field description of an <strong>in</strong>teract<strong>in</strong>g BEC-0.015-0.02-0.025-0.03-0.035-0.04Numerical resultsFirst-order correction-0.04510 5 10 6 10 7Figure 4.9: A relative shift <strong>in</strong> the condensation temperature δT c as a function of thenumber of particles <strong>in</strong> the trap N for ω = 100 Hz.Nwe approach the BEC phase transition from the gas phase. We stress that the sameresult can be obta<strong>in</strong>ed by consider<strong>in</strong>g condensate phase <strong>in</strong> the vic<strong>in</strong>ity of the phasetransition with<strong>in</strong> the TFSC model. It is important to note that we only obta<strong>in</strong>the <strong>in</strong>teraction-<strong>in</strong>duced shift <strong>in</strong> the condensation temperature. F<strong>in</strong>ite-size effects[91, 92] are beyond our consideration, because we rely on the TF and semiclassicalapproximation. A comparison of the analytical expression (4.34), which gives a firstordercorrection to T c <strong>in</strong> the <strong>in</strong>teraction strength, and numerical results based onEqs. (4.26) and (4.27), is shown <strong>in</strong> Fig. 4.9. In this graph, δT c is given as a functionof the number of particles <strong>in</strong> the trap. We observe a reasonable agreement of theTFSC result (4.34) with the numerical data.4.3 Experimental assessment of different modelsDensity profile of a BEC cloud is experimentally measurable quantity, us<strong>in</strong>g eitherthe absorption TOF imag<strong>in</strong>g or <strong>in</strong>-situ techniques described <strong>in</strong> Chapter 1. Thevalues of other physical quantities are then extracted from the experimental data <strong>in</strong>a roundabout way.Early experimental efforts [93, 94] to measure the occupation of the ground stateas a function of the temperature have usually employed the absorption TOF imag<strong>in</strong>g.Thermometry was then performed by fitt<strong>in</strong>g the semiclassical result (1.17)100


4. Mean-field description of an <strong>in</strong>teract<strong>in</strong>g BECto the w<strong>in</strong>gs of the measured velocity distribution of a thermal cloud. The condensatecomponent was identified as a characteristic narrow central peak, and thecondensate fraction was obta<strong>in</strong>ed by fitt<strong>in</strong>g either an additional Gaussian profile ora parabolic TF profile to that part of the distribution, <strong>in</strong> accordance with the previouslydescribed almost-ideal model. These early experimental studies were only ableto validate that the phase diagram of the system is quite accurately described bythe non<strong>in</strong>teract<strong>in</strong>g result from Eq. (1.10). The <strong>in</strong>herent statistical and systematicerrors prevented real quantitative tests of the mean-field models.The first clear experimental demonstration, which measured values of condensatefractions different from the correspond<strong>in</strong>g ideal values of Eq. (1.10), was reported<strong>in</strong> Ref. [89]. The experimental procedure <strong>in</strong>cluded TOF absorption imag<strong>in</strong>g andan additional technique, a coherent Brag scatter<strong>in</strong>g. The advantage of the lattertechnique is its ability to provide a complete spatial separation of the condensateand thermal cloud dur<strong>in</strong>g the TOF expansion. The condensate fraction was determ<strong>in</strong>edwith a high certa<strong>in</strong>ty by count<strong>in</strong>g the scattered atoms. However, once aga<strong>in</strong> aballistic expansion of the thermal cloud was assumed, and the temperature was determ<strong>in</strong>edby fitt<strong>in</strong>g the semiclassical density profile of the expanded non<strong>in</strong>teract<strong>in</strong>ggas (1.13) to the measured density profile of the thermal cloud. Despite this, the obta<strong>in</strong>edphase diagram clearly demonstrated a reduced condensate fraction comparedto the non<strong>in</strong>teract<strong>in</strong>g result (1.10). A good agreement of experimental values withthe results obta<strong>in</strong>ed with<strong>in</strong> the semi-ideal model was found, while the agreement waseven better with the predictions of the TFSC model. The ma<strong>in</strong> focus of Ref. [89]was a precise measurement of the <strong>in</strong>teraction-<strong>in</strong>duced shift of the condensation temperatureand its comparison with Eq. (4.34). One of the ma<strong>in</strong> conclusions was thatimproved models are necessary for a good quantitative description of the expansionof f<strong>in</strong>ite-temperature BECs and reliable analysis of experimental data.In order to avoid <strong>in</strong>tricacies of the accurate model<strong>in</strong>g of the TOF expansion,Ref. [24] has exploited the <strong>in</strong>-situ phase-contrast imag<strong>in</strong>g to <strong>in</strong>vestigate the BECphase diagram experimentally. The values of the chemical potential µ and thetemperature T were determ<strong>in</strong>ed by fitt<strong>in</strong>g to functions that stem from the <strong>in</strong>teract<strong>in</strong>gf<strong>in</strong>ite-temperature models to the <strong>in</strong>-situ measured data. Most reliable fits wereobta<strong>in</strong>ed <strong>in</strong> the case of the Popov model [78], which is very similar to the presentedTFSC model. We stress that authors of Ref. [24] have used the consistent procedure:the same mean-field model was used for the extraction of all relevant quantities (µ, T,N 0 , N) from the experimental results, as well as for the analysis of the phase diagram.101


4. Mean-field description of an <strong>in</strong>teract<strong>in</strong>g BECAs one of the chief outcomes of that study, the mean-field <strong>in</strong>teraction-<strong>in</strong>duced shift<strong>in</strong> the condensation temperature, Eq. (4.34), was experimentally confirmed.F<strong>in</strong>ally, a new series of experiments [80, 81] <strong>in</strong> 2011 addressed the same topic byutiliz<strong>in</strong>g the Feshbach resonance technique of 39 K <strong>in</strong> comb<strong>in</strong>ation with absorptionTOF imag<strong>in</strong>g. Several advantageous aspects of a Feshbach resonance were used.First of all, a wide range of <strong>in</strong>teraction strengths was covered. Also, referencemeasurements for the same number of atoms and different <strong>in</strong>teraction strengthswere performed. Furthermore, a ballistic expansion was carefully eng<strong>in</strong>eered bytun<strong>in</strong>g the value of the scatter<strong>in</strong>g length to zero dur<strong>in</strong>g the TOF expansion. Forthe weakly <strong>in</strong>teract<strong>in</strong>g gas, the mean-field correction result (4.34) was found to be<strong>in</strong> good agreement with the experimental data. In addition, for the case of stronger<strong>in</strong>teractions, beyond-mean-field effects were clearly observed for the first time.4.4 Conclusions and outlookIn this Chapter we have reviewed the mean-field description of harmonically trappedBECs. In the zero-temperature limit we have presented widely used Gross-Pitaevskiiequation, which we will further exploit <strong>in</strong> Chapter 5. The relevance of the f<strong>in</strong>itetemperaturemean-field models stems from the fact that they are regularly used <strong>in</strong>the <strong>in</strong>terpretation of experimental data and extraction of density profiles for thecondensate and for the thermal cloud. We have considered four different f<strong>in</strong>itetemperaturemean-field models. By compar<strong>in</strong>g the predictions of all the consideredmodels, we have concluded that the TFSC model provides the most appropriateand consistent choice from the practical po<strong>in</strong>t of view. It is quite easy for numericalimplementation, and is capable to reasonably describe the BEC phase transition.With<strong>in</strong> this approach we have derived the <strong>in</strong>teraction-<strong>in</strong>duced shift of the condensationtemperature. The model, however, has a major drawback <strong>in</strong> that it leads to theunphysical discont<strong>in</strong>uities <strong>in</strong> the density profiles. The semi-ideal and GPSC modelyield cont<strong>in</strong>uous density profiles, but do not capture thermodynamic properties ofthe system properly. Some of the mentioned issues are even more serious <strong>in</strong> thecase of a more complex trap geometry or <strong>in</strong> the rotat<strong>in</strong>g systems [95]. The resultsobta<strong>in</strong>ed so far po<strong>in</strong>t to the necessity of us<strong>in</strong>g improved semiclassical approximationsfor the thermal atoms, which are better suited to the quantum-mechanicaldescription of the ground-state. Now that experiments have succeeded <strong>in</strong> achiev<strong>in</strong>gthe clear observation of beyond-mean-field effects, the improvements <strong>in</strong> widely used102


approximations become a demand.4. Mean-field description of an <strong>in</strong>teract<strong>in</strong>g BEC103


Chapter 5Nonl<strong>in</strong>ear BEC dynamics by harmonicmodulation of s-wave scatter<strong>in</strong>g lengthA thorough knowledge of the excitation spectra of a quantum system is quiteimportant s<strong>in</strong>ce the properties of its excitations characterize the phase of the matter<strong>in</strong> a very precise way. The excitation spectra capture <strong>in</strong>formation on the system’scorrelations and def<strong>in</strong>e the response of the system to external perturbations. Experimentally,they are determ<strong>in</strong>ed by expos<strong>in</strong>g the system to a weak perturbationand by measur<strong>in</strong>g the <strong>in</strong>duced dynamics. Thus, the dynamics of the system and itsexcitation spectra are closely related concepts.First microscopic derivation of the excitation spectrum of a superfluid, given byEq. (1.28), was published <strong>in</strong> the sem<strong>in</strong>al paper of Bogoliubov <strong>in</strong> 1947 [30]. The obta<strong>in</strong>edspectrum conta<strong>in</strong>s a collective phonon mode <strong>in</strong> the limit of a vanish<strong>in</strong>g wavenumber, and s<strong>in</strong>gle-particle excitations <strong>in</strong> the limit of a large wave number. Theorig<strong>in</strong>al aim of Bogoliubov was the explanation of the superfluidity <strong>in</strong> 4 He and thecollective motion of bosonic particles was identified as the underly<strong>in</strong>g mechanismthat leads to the phenomenon. Yet, the Bogoliubov approach assumes weak <strong>in</strong>teractions,while 4 He is a strongly <strong>in</strong>teract<strong>in</strong>g system and the agreement between thetheoretical and experimental results is only qualitative. The dilute vapors of alkaliatoms were the first proper experimental realization of a weakly <strong>in</strong>teract<strong>in</strong>g bosonicsystem that allowed quantitative tests of the theoretical concepts <strong>in</strong>troduced <strong>in</strong> theearly papers of Landau [96] and Bogoliubov [30]. For this reason, the experimentsstudy<strong>in</strong>g the collective bosonic modes were <strong>in</strong>itiated soon after the first achievementof an atomic BEC.A basic underly<strong>in</strong>g idea of the early experiments was a clear demonstrationof the dist<strong>in</strong>ction between the collective response of the condensed cloud and theresponse of thermal atoms to external perturbations. A common scenario <strong>in</strong> themajority of experiments is to start with the ground-state condensate at very lowtemperature, where the condensate depletion is negligible, and to <strong>in</strong>duce collectiveoscillation modes by a temporary modulation of the external trapp<strong>in</strong>g potential104


5. BEC excitation by modulation of scatter<strong>in</strong>g length[97, 98, 99, 100, 101]. For <strong>in</strong>stance, <strong>in</strong> the early experiment reported <strong>in</strong> Ref. [97]collective modes were excited by apply<strong>in</strong>g a small time-dependent perturbation of agiven frequency to the transversal component of the trap-potential and the real-timedynamics <strong>in</strong> terms of shape oscillations of the condensate was observed. Based onthese measurements, two low-ly<strong>in</strong>g eigenmodes of different symmetry were identified.The same procedure was repeated at higher temperatures, when the condensateis not present, and, as expected, the thermal cloud has produced only a responsecorrespond<strong>in</strong>g to the excitations of a normal non<strong>in</strong>teract<strong>in</strong>g gas. In the experimentalstudy published <strong>in</strong> the same year Ref. [98], the condensate was excited by thetime-dependent modulation of the trapp<strong>in</strong>g potential, which additionally <strong>in</strong>cludeda spatial displacement of the potential m<strong>in</strong>imum. In this case, shape oscillationscoupled with the center of mass motion were observed. A typical experimental observationis given <strong>in</strong> Fig. 5.1. In order to make a detailed comparison with theoreticalmodels, the eigenmode frequencies were measured for different numbers of trappedatoms and for different trap strengths.Figure 5.1: An experimental measurement of the collective BEC modes. A figureobta<strong>in</strong>ed by Stamper-Kurn and Ketterle, taken from Ref. [5]. The time evolutionof a BEC cloud is presented by a series of density profiles. Shape oscillations arecoupled with the center of mass motion. The field of view <strong>in</strong> the vertical directionis about 620 µm and time step is 5 ms per frame.The orig<strong>in</strong>al derivation of the Bogoliubov excitation spectrum was performed fora translationally <strong>in</strong>variant system. However, <strong>in</strong> cold-atom experiments, bosons arealways conf<strong>in</strong>ed by the external trap and we have to deal with an <strong>in</strong>homogeneoussystem. In order to be able to keep up with the experimental advances, new theoreticalapproaches for the description of collective modes of a trapped system were105


5. BEC excitation by modulation of scatter<strong>in</strong>g lengthdeveloped. In the mean-field framework, l<strong>in</strong>earization of the GP equation (4.15)around the ground-state [102] is analogous to the Bogoliubov approach for the <strong>in</strong>homogeneouscase. Of course, the procedure is not analytically tractable and has to beimplemented numerically. Most widespreadly used analytical approximative results,which also exploit GP equation as a start<strong>in</strong>g po<strong>in</strong>t, are Str<strong>in</strong>gari’s results obta<strong>in</strong>ed <strong>in</strong>the Thomas-Fermi limit [103], and variational results presented <strong>in</strong> Refs. [104, 105].A very good agreement between the experimental values of the frequencies of lowly<strong>in</strong>gmodes and theoretical results based on the l<strong>in</strong>earized mean-field equations wasestablished [102, 103, 104] and validity of the mean-field description, given by theGP equation (4.15), was fully confirmed.All results mentioned until now have assumed small-amplitude oscillations andthus rely on a l<strong>in</strong>earized regime of the <strong>in</strong>itial GP equation. Once this regime wasexplored to some extent, more complex, physically content-rich, and more <strong>in</strong>terest<strong>in</strong>gdynamical features became the focus of experimental studies. We will mention themost prom<strong>in</strong>ent examples. As a first notable example of nonl<strong>in</strong>ear collective BECexcitations we mention localized solutions, usually called solitons [19, 85, 86]. Asalready po<strong>in</strong>ted out, the ma<strong>in</strong> equation that describes the dynamics of a BEC atT = 0, the GP equation (4.15), is a nonl<strong>in</strong>ear Schröd<strong>in</strong>ger equation. This typeof equations is studied extensively <strong>in</strong> the field of nonl<strong>in</strong>ear optics and from thiscontext it is well-known that it admits nonl<strong>in</strong>ear localized solutions. Several recentexperiments [106, 107] have studied creation and <strong>in</strong>teraction of solitons <strong>in</strong> the atomicBEC. Another subject of wide <strong>in</strong>terest is pattern formation <strong>in</strong> a driven system, <strong>in</strong>particular formation of Faraday patterns [108, 109, 110]. An observation of this typeof dynamics <strong>in</strong> a BEC was given <strong>in</strong> Ref. [111], where a density wave <strong>in</strong> the axialdirection was produced by a strong modulation of the strength of the radial trapp<strong>in</strong>gpotential. A further experimental research <strong>in</strong>cludes study of a quantum turbulentregime <strong>in</strong> a BEC [112] by a comb<strong>in</strong>ation of rotation, strong modulation of a trapstrength and a trap displacement. In essence, quantum turbulence is a superfluidturbulence characterized by the presence of tangled vortices. It was <strong>in</strong>itially studied<strong>in</strong> the superfluid Helium, but now it can be studied <strong>in</strong> a more controlled way <strong>in</strong> aBEC setup, as shown <strong>in</strong> Ref. [112].In addition to already mentioned research avenues, an excit<strong>in</strong>g possibility forreach<strong>in</strong>g a non-trivial nonl<strong>in</strong>ear dynamical regime <strong>in</strong> a BEC cloud is given by areal-time tun<strong>in</strong>g of the <strong>in</strong>teraction strength via a Feshbach resonance mechanism<strong>in</strong>troduced <strong>in</strong> Chapter 1. Harmonic modulation of the s-wave scatter<strong>in</strong>g length as106


5. BEC excitation by modulation of scatter<strong>in</strong>g lengtha method for excitation of collective oscillation modes was proposed <strong>in</strong> Refs. [113,114, 115, 116], but it was experimentally realized only recently <strong>in</strong> Ref. [15]. In themean-field approximation at T = 0, the time-dependent <strong>in</strong>teraction leads to a timedependentnonl<strong>in</strong>earity g(t) <strong>in</strong> the GP equation. Depend<strong>in</strong>g on the closeness of theexternal modulation frequency Ω to one of condensate’s eigenmodes, a qualitativelydifferent dynamical behaviors emerge. In the non-resonant case, we have smallamplitudeoscillations of the condensate size around the equilibrium widths, and weare <strong>in</strong> the regime of l<strong>in</strong>ear response. However, as Ω approaches an eigenmode, weexpect a resonant behavior which is characterized by large amplitude oscillations.In this case it is clear that a l<strong>in</strong>ear response analysis does not provide a qualitativelygood description of the system dynamics.Motivated by the experimental study, described <strong>in</strong> Ref. [15], <strong>in</strong> this Chapter weconsider dynamical features <strong>in</strong>duced by harmonic modulation of the s-wave scatter<strong>in</strong>glength. Our study is a step beyond the l<strong>in</strong>ear regime, toward the resonantbehavior, and it is suited for the parametric region where low-ly<strong>in</strong>g collective modescan still be def<strong>in</strong>ed as <strong>in</strong> the l<strong>in</strong>ear regime, but their properties are modified bynonl<strong>in</strong>ear effects. Obta<strong>in</strong>ed results are relevant for the proper <strong>in</strong>terpretation ofexperimental data, and for understand<strong>in</strong>g of near-resonant properties of nonl<strong>in</strong>earsystems.In the follow<strong>in</strong>g, we first review variational description of low-ly<strong>in</strong>g modes <strong>in</strong> thel<strong>in</strong>ear regime. Then, we turn to the recent experiment, published <strong>in</strong> Ref. [15], thathas achieved harmonic modulation of the s-wave scatter<strong>in</strong>g length and briefly expla<strong>in</strong>experimental procedure and results. F<strong>in</strong>ally, we study the nonl<strong>in</strong>ear dynamicalregime <strong>in</strong>duced by harmonic modulation of the s-wave scatter<strong>in</strong>g length, first for aspherically symmetric BEC, and afterwards for an axially-symmetric BEC. In bothcases, we obta<strong>in</strong> excitation spectra as Fourier transforms of the time-dependent condensatesizes and from here we identify nonl<strong>in</strong>ear features. In addition, we developperturbation theory based on the Po<strong>in</strong>caré-L<strong>in</strong>dstedt method which successfully expla<strong>in</strong>sthe observed nonl<strong>in</strong>ear effects.5.1 Variational description of low-ly<strong>in</strong>g modesOur analytical method of choice for study<strong>in</strong>g nonl<strong>in</strong>ear BEC dynamics is variationalapproach <strong>in</strong>troduced <strong>in</strong> Refs. [104, 105]. For completeness and for <strong>in</strong>structivereasons, we first present the method and ma<strong>in</strong> results on low-ly<strong>in</strong>g collective BEC107


5. BEC excitation by modulation of scatter<strong>in</strong>g lengthmodes obta<strong>in</strong>ed previously <strong>in</strong> the l<strong>in</strong>ear regime. As expla<strong>in</strong>ed <strong>in</strong> Chapter 4, thetime-dependent GP equation can be obta<strong>in</strong>ed by extremiz<strong>in</strong>g the functional (4.16)with respect to ψ(⃗r, t). In the core of the variational description is the idea to plug<strong>in</strong> an Ansatz for the condensate wave function <strong>in</strong>to Eq. (4.16) and to derive the correspond<strong>in</strong>gEuler-Lagrange equations of the system with respect to the parameterspresent <strong>in</strong> the Ansatz. In this way, <strong>in</strong>stead of the partial differential GP equation,we reduce the description of a system to ord<strong>in</strong>ary differential equations, which is farsimpler. Naturally, this presents an approximation, and careful exam<strong>in</strong>ation of itsvalidity and associated errors is necessary.To this end, we closely follow derivations presented orig<strong>in</strong>ally <strong>in</strong> Refs. [104, 105].We assume that the condensate wave function has the same Gaussian form <strong>in</strong> the<strong>in</strong>teract<strong>in</strong>g case as <strong>in</strong> the non<strong>in</strong>teract<strong>in</strong>g one, just with renormalized parameters.Thus, we use a time-dependent variational method based on a Gaussian Ansatz,which for the anisotropic harmonic trap (1.19) readsψ G (x, y, z, t) = N(t)∏σ=x,y,z[exp − 1 ](σ − σ 0 (t)) 2+ i σϕ2 u σ (t) 2 σ (t) + iσ 2 φ σ (t) , (5.1)where N(t) = π −3 4u x (t) −1 2u y (t) −1 2u z (t) −1 2 is a time-dependent normalization, whileu σ (t), φ σ (t), σ 0 and ϕ σ are variational parameters. For convenience, throughoutthis Chapter, we normalize the wave function to unity and for consistency we <strong>in</strong>cludethe total number of atoms <strong>in</strong>to the correspond<strong>in</strong>g <strong>in</strong>teraction strength. Thus,we modify the notation <strong>in</strong>troduced earlier by perform<strong>in</strong>g the follow<strong>in</strong>g transformation:ψ(⃗r, t) → ψ(⃗r, t)/ √ N, g → g × N. The <strong>in</strong>troduced variational parametershave straightforward <strong>in</strong>terpretation: u σ (t) parameters correspond to the condensatewidths <strong>in</strong> different directions and are roughly proportional to the root-mean-squarewidths of the exact condensate wave function ψ(x, y, z, t); φ σ (t) and ϕ σ (t) parametersrepresent the correspond<strong>in</strong>g phases of the wave function and are essential forthe proper description of dynamical features; a possible center-of-mass motion iscaptured by the parameters σ 0 .Follow<strong>in</strong>g Ref. [104], we <strong>in</strong>sert Ansatz (5.1) <strong>in</strong>to the Lagrangian (4.16) yield<strong>in</strong>gthe GP equation, and extremize it with respect to variational parameters. All thedetails of the derivation are given <strong>in</strong> Appendix B and here we give only a briefexplanation. By extremiz<strong>in</strong>g the functional, we first obta<strong>in</strong> a coupled system ofdifferential equations of the first order for all variational parameters. The equations108


5. BEC excitation by modulation of scatter<strong>in</strong>g lengthfor the phases φ σ and ϕ σ can be solved explicitly <strong>in</strong> terms of the widths u σ andthe center of mass coord<strong>in</strong>ates σ 0 . In this way we obta<strong>in</strong> two sets of ord<strong>in</strong>arysecond-order differential equations that govern the condensate dynamics. A centerof-massmotion is decoupled from the shape oscillations and is determ<strong>in</strong>ed by simpleharmonic oscillator equations, which are <strong>in</strong>dependent of <strong>in</strong>teratomic <strong>in</strong>teractions:¨σ 0 (t) + λ 2 σσ 0 (t) = 0, σ ∈ {x, y, z} . (5.2)On the other hand, the widths of the condensate exhibit non-trivial dynamics givenby a set of coupled nonl<strong>in</strong>ear differential equations:ü σ (t) + λ 2 σu σ (t) − 1u σ (t) − P= 0, σ ∈ {x, y, z} . (5.3)3 u σ (t)u x (t)u y (t)u z (t)In the previous equations and from now on, we use dimensionless notation: wechoose a convenient frequency scale ω (for example, the external trap frequency <strong>in</strong>one of the spatial directions) and express all lengths <strong>in</strong> the units of the characteristicharmonic oscillator length l = √ /Mω, time <strong>in</strong> units of ω −1 and external frequencies<strong>in</strong> units of ω: λ σ = ω σ /ω, σ ∈ {x, y, z}. The dimensionless <strong>in</strong>teraction parameter Pis given by P = g/((2π) 3/2 ωl 3 ) = √ 2/πNa/l.In this approach, the <strong>in</strong>itial partial differential equation (4.15) is approximatedwith the two sets of ord<strong>in</strong>ary differential equations, given by Eqs. (5.2) and (5.3),which allow analytical considerations. The first properties of the condensate thatcan be calculated are the equilibrium widths u x0 , u y0 and u z0 . They are found bysolv<strong>in</strong>g an algebraic system of equations:λ 2 σu σ0 − 1 P− = 0, σ ∈ {x, y, z} . (5.4)u σ0 u σ0 u x0 u y0 u z0The equilibrium widths represent stationary solutions of Eqs. (5.3).Now we turn to the calculation of the frequencies of low-ly<strong>in</strong>g modes. To beg<strong>in</strong>with, from Eqs.(5.2) we read off frequencies that correspond to the center-of-massmotion. These are dipole modes and their frequencies are equal to the external trapfrequencies (for the case of a harmonic trap). Most often, this type of excitationsis created by shift<strong>in</strong>g the trap <strong>in</strong> space. Actually, the well established experimentalprocedure for the precise determ<strong>in</strong>ation of the trap parameters is based on themeasurement of the dipole mode frequencies [21].109


5. BEC excitation by modulation of scatter<strong>in</strong>g lengthNext, by l<strong>in</strong>eariz<strong>in</strong>g the Eqs. (5.3) around the equilibrium widths (5.4), we canobta<strong>in</strong> <strong>in</strong>formation on collective modes related to the BEC shape oscillations. Wewill consider experimentally relevant case of an axially symmetric trap, such thatλ x = λ y = 1, u x0 = u y0 ≡ u ρ0 . Due to the axial symmetry of the consideredsystem, the projection of the angular momentum along the z-axis, L z , is a goodquantum number and we use it for the classification of the modes. By solv<strong>in</strong>gthe correspond<strong>in</strong>g eigenproblem, we f<strong>in</strong>d three different modes depicted <strong>in</strong> Fig. 5.2.Accord<strong>in</strong>g to their symmetry properties, we designate the modes as the |L z | = 2quadrupole mode, the L z = 0 quadrupole mode, and the breath<strong>in</strong>g mode (whichalso corresponds to L z = 0). Eigenvalues of the l<strong>in</strong>earized system of equations yieldthe frequencies of the collective modes. Frequencies of the L z = 0 quadrupole modeω Q0 , and the breath<strong>in</strong>g mode ω B0 are given by:ω B0,Q0 = √ 2[ (1 + λ 2 z − P4u 2 ρ0 u3 z0)√ (± 1 − λ 2 z +P4u 2 ρ0 u3 z0while the frequency of the |L z | = 2 quadrupole mode is given by:ω |Lz|=2Q0=) 2 ( )]1/2 2 P+ 8,4u 3 ρ0 u2 z0(5.5)√P4 − 2 . (5.6)u 4 ρ0u z0As shown <strong>in</strong> Fig. 5.2, the |L z | = 2 quadrupole mode is characterized by out-ofphaseoscillations <strong>in</strong> x and y directions, the |L z | = 0 quadrupole mode exhibitsout-of-phase radial and axial oscillations, while <strong>in</strong>-phase radial and axial oscillationscorrespond to the breath<strong>in</strong>g mode.Figure 5.2: A schematic illustration of the condensate eigenmodes: |L z | = 2quadrupole mode (left), |L z | = 0 quadrupole mode (middle) and |L z | = 0 breath<strong>in</strong>gmode (right).110


5. BEC excitation by modulation of scatter<strong>in</strong>g lengthThe ma<strong>in</strong> result obta<strong>in</strong>ed by us<strong>in</strong>g the Gaussian approximation is an analyticalestimate for the frequencies of the low-ly<strong>in</strong>g collective modes, expressed by Eqs. (5.5)and (5.6) [104, 105]. We emphasize that, although based on the Gaussian Ansatz,the variational approximation reproduces exactly the frequencies of collective modesnot only for the weakly <strong>in</strong>teract<strong>in</strong>g BEC, but also for the strongly <strong>in</strong>teract<strong>in</strong>g BEC<strong>in</strong> the Thomas-Fermi regime [103, 104]. Therefore, it represents a solid analyticaldescription of BEC dynamics. Most importantly, a reasonable quantitative agreementwas obta<strong>in</strong>ed between the l<strong>in</strong>ear response theoretical results (5.5) and (5.6)and experimental results for BEC excited us<strong>in</strong>g the trap modulation. In general, adetailed experimental <strong>in</strong>formation on collective modes allows us to test our theoreticalunderstand<strong>in</strong>g of the properties of an atomic BEC. The essential merit of test<strong>in</strong>gtheoretical predictions us<strong>in</strong>g collective oscillation modes stems from the possibilityto measure frequencies of collective modes with a high accuracy on the order of lessthan 1% [100, 99, 101].We po<strong>in</strong>t out that the early theoretical studies of collective modes of BEC[102, 103, 104] focused on the exploration of dynamical properties <strong>in</strong> the l<strong>in</strong>earregime of small amplitude oscillations. Certa<strong>in</strong> nonl<strong>in</strong>ear aspects of condensate dynamics<strong>in</strong>duced by a trap modulation are given <strong>in</strong> Refs. [117, 118, 119], whereastwo-component BECs are dealt with <strong>in</strong> Refs. [120, 121]. In the next section we turnto the recent experimental study <strong>in</strong> which nonl<strong>in</strong>ear effects arise from the modulationof the <strong>in</strong>teraction strength and study how nonl<strong>in</strong>ear dynamical features arereflected on the properties of the excitation spectrum.5.2 Harmonic modulation of the s-wave scatter<strong>in</strong>g length:experimentDetails of the Feshbach resonance of a 7 Li BEC were explored by the R. Hulet’sgroup from the Rice University <strong>in</strong> Ref. [122], and an extreme tunability of <strong>in</strong>teractionswas experimentally demonstrated. In the mentioned experiment, atoms weretrapped by the optical trap, while the bias magnetic field was used for tun<strong>in</strong>g thescatter<strong>in</strong>g length via the Feshbach resonance. For the range of values of magneticfield B, a ground-state condensate was produced and the correspond<strong>in</strong>g densityprofiles were observed. By measur<strong>in</strong>g the width of the condensate distribution andcompar<strong>in</strong>g these values with the correspond<strong>in</strong>g numerical data based on the GPequation (4.15), <strong>in</strong>formation on the scatter<strong>in</strong>g length spann<strong>in</strong>g seven orders of mag-111


5. BEC excitation by modulation of scatter<strong>in</strong>g lengthnitude was extracted, as can be seen <strong>in</strong> Fig. 5.3. In this way, precise values of theparameters of the Feshbach resonance (1.23) were obta<strong>in</strong>ed:a BG = −24.5a 0 , ∆ B = 19.23(3) mT, B ∞ = 73.68(2) mT,where a 0 represents the Bohr radius.10 6 Magnetic Field (mT)Scatter<strong>in</strong>g Length (a 0)10 410 210 00.60.40.20.0-0.254.0 54.2 54.4 54.6 54.8 55.010 -255.0 60.0 65.0 70.0Figure 5.3: Feshbach resonance of a 7 Li BEC: Dependence of the scatter<strong>in</strong>g lengthon the external magnetic field. Results are taken from Ref. [122].In the follow-up paper, Ref. [15], the same group <strong>in</strong> collaboration with the groupof V. Bagnato from Sao Pãulo University, has experimentally realized an alternativeway of a 7 Li condensate excitation, based on us<strong>in</strong>g the Feshbach resonance. Byperiodically chang<strong>in</strong>g an external magnetic field as B(t) = B av + δB cos Ωt, theywere able to obta<strong>in</strong> harmonic modulation of the s-wave scatter<strong>in</strong>g length <strong>in</strong> theforma(t) ≃ a av + δ a cos Ωt , (5.7)where a av = a(B av ) ≈ 3a 0 > δ a = − a BG∆ B δB(B av−B ∞) 2 ≈ 2a 0 . In this way, a time-dependent<strong>in</strong>teraction among atoms was realized, which is expressed <strong>in</strong> terms of the dimensionlessparameter P(t) asP(t) = P + Q cosΩt , (5.8)112


5. BEC excitation by modulation of scatter<strong>in</strong>g lengthwhere P = √ 2/πNa av /l denotes the average <strong>in</strong>teraction strength, Q = √ 2/πNδ a /lis a modulation amplitude, and Ω represents the modulation or driv<strong>in</strong>g frequency.The dimensionless experimental parameters from Ref. [15] have the valuesP = 15, Q = 10, λ z = 0.021 , (5.9)correspond<strong>in</strong>g to the highly elongated trap with the modulated but always positive(repell<strong>in</strong>g) <strong>in</strong>teraction. In the experiment, the oscillations of the condensate sizewere observed by <strong>in</strong>-situ phase-contrast imag<strong>in</strong>g, presented <strong>in</strong> Fig. 5.4.Figure 5.4: Oscillations of the BEC cloud, presented by a series of density profilestaken at equidistant time step of 15 ms. Results are taken from Ref. [15].The <strong>in</strong>terpretation of the experimental data was based on the analytical resultsfor the frequencies of the low-ly<strong>in</strong>g collective modes obta<strong>in</strong>ed from the l<strong>in</strong>earizedform of the Gaussian approximation. For experimental data, Eq. (5.5) yields thefollow<strong>in</strong>g values for the frequencies of the quadrupole and the breath<strong>in</strong>g mode:ω Q0 = 0.035375 , ω B0 = 2.00002 . (5.10)The external trap was stationary, thus prevent<strong>in</strong>g excitations of the dipole (Kohn)mode, correspond<strong>in</strong>g to the center-of-mass motion. For the specific set of experimentalparameters basically only the quadrupole oscillation mode was excited <strong>in</strong> thisway and resonances located at the quadrupole frequency and its second harmonicwere observed.There are several advantages of such an experimental scheme: for <strong>in</strong>stance, <strong>in</strong>future experiments with multi-species BEC, a s<strong>in</strong>gle component could be <strong>in</strong>dividuallyexcited <strong>in</strong> this way, while the excitation of other components would occur only<strong>in</strong>directly.113


5. BEC excitation by modulation of scatter<strong>in</strong>g lengthHowever, due to the nonl<strong>in</strong>ear form of the underly<strong>in</strong>g GP equation, we expectnonl<strong>in</strong>earity-<strong>in</strong>duced shifts <strong>in</strong> the frequencies of low-ly<strong>in</strong>g modes compared to thevalues obta<strong>in</strong>ed from Eq. (5.5), calculated us<strong>in</strong>g the l<strong>in</strong>ear stability analysis. Inparticular, <strong>in</strong> the case of a close match<strong>in</strong>g of the driv<strong>in</strong>g frequency Ω and one ofthe BEC eigenmodes, we expect resonances, i.e. large amplitude oscillations of thecondensate size. When this happens, the role of the nonl<strong>in</strong>ear terms becomes crucialand nonl<strong>in</strong>ear phenomena become dom<strong>in</strong>ant. Furthermore, we emphasize thatoscillations with very small amplitudes, which occur <strong>in</strong> the l<strong>in</strong>ear regime, are experimentallyhard to observe. On the other hand, very large amplitude oscillationslead to a fragmentation of the condensate [15, 123]. Thus, the case <strong>in</strong> between is ofthe ma<strong>in</strong> experimental <strong>in</strong>terest and represents our ma<strong>in</strong> objective, as we discuss <strong>in</strong>the next section.5.3 Harmonic modulation of the s-wave scatter<strong>in</strong>g length:theoretical frameworkTo study nonl<strong>in</strong>ear BEC dynamics, we use an approach that is complementary tothe recent theoretical considerations [113, 114, 115, 116] of a BEC with harmonicallymodulated <strong>in</strong>teraction. In Ref. [114] the real-time dynamics of a sphericallysymmetric BEC was numerically studied and analytically expla<strong>in</strong>ed us<strong>in</strong>g the resonantBogoliubov-Mitropolsky method [124]. On the other hand, <strong>in</strong> our approach <strong>in</strong>order to discern <strong>in</strong>duced dynamical features, we look directly at the excitation spectrumobta<strong>in</strong>ed from a Fourier transform of the time-dependent condensate width.From this type of numerical analysis we f<strong>in</strong>d characteristic nonl<strong>in</strong>ear properties:higher harmonic generation, nonl<strong>in</strong>ear mode coupl<strong>in</strong>g, and significant shifts <strong>in</strong> thefrequencies of collective modes with respect to their l<strong>in</strong>ear response counterparts.In addition, we work out an analytic perturbative theory with respect to the modulationamplitude, capable of captur<strong>in</strong>g many of the mentioned nonl<strong>in</strong>ear effectsobta<strong>in</strong>ed numerically.Nonl<strong>in</strong>earity-<strong>in</strong>duced frequency shifts were considered previously <strong>in</strong> Ref. [117] forthe case of bosonic collective modes excited by modulation of the trapp<strong>in</strong>g potential,and also <strong>in</strong> Ref. [125] for the case of a superfluid Fermi gas <strong>in</strong> the BCS-BEC crossover.Our analytical approach is based on the Po<strong>in</strong>caré-L<strong>in</strong>dstedt method [126, 127, 128,124], <strong>in</strong> the same spirit as presented <strong>in</strong> Refs. [117, 125, 126]. However, the harmonicmodulation of the <strong>in</strong>teraction strength <strong>in</strong>troduces additional features that require a114


separate treatment.5. BEC excitation by modulation of scatter<strong>in</strong>g lengthIn Ref. [116] it was predicted that a harmonic modulation of the scatter<strong>in</strong>g lengthleads to the creation of Faraday patterns <strong>in</strong> BEC, i.e. density waves. Up to now,Faraday patterns have been experimentally <strong>in</strong>duced by harmonic modulation of thetransverse conf<strong>in</strong>ement strength [111], which is studied analytically and numerically<strong>in</strong> Ref. [129]. Here we focus only on the nonl<strong>in</strong>ear properties of low-ly<strong>in</strong>g collectivemodes and do not consider possible excitations of Faraday patterns.In order to obta<strong>in</strong> analytical <strong>in</strong>sight <strong>in</strong>to the condensate dynamics <strong>in</strong>duced bythe harmonic modulation of the s-wave scatter<strong>in</strong>g length described by Eq. (5.7), weuse the Gaussian variational approximation approximation. We consider an axiallysymmetric BEC, (λ x = λ y = 1, u x = u y ≡ u ρ ), excited by modulation of the<strong>in</strong>teraction strength, which preserves the axial symmetry of the condensate dur<strong>in</strong>gits time evolution. For this reason, we use a simplified axially-symmetric form ofEqs. (5.3):ü ρ (t) + u ρ (t) − 1u ρ (t) 3 −P(t)u ρ (t) 3 u z (t)= 0 , (5.11)ü z (t) + λ 2 zu z (t) − 1u z (t) − P(t)= 0 , (5.12)3 u ρ (t) 2 u z (t) 2which we will refer to as a Gaussian approximation.To estimate the accuracy of the Gaussian approximation for describ<strong>in</strong>g the dynamics<strong>in</strong>duced by the harmonic modulation of the <strong>in</strong>teraction strength, we compareits solution with an exact numerical solution of the GP equation. In Fig. 5.5, weplot the result<strong>in</strong>g time-dependent axial and radial condensate widths ρ rms (t) andz rms (t), calculated as root mean square valuesρ rms (t) =z rms (t) =√√2π2π∫ ∞−∞∫ ∞−∞dzdz∫ ∞0∫ ∞0ρ dρ |ψ(ρ, z, t)| 2 ρ 2 , (5.13)ρ dρ |ψ(ρ, z, t)| 2 z 2 , (5.14)of the solution of the GP equation, as well as numerical solutions of Eqs. (5.11)and (5.12). We assume that <strong>in</strong>itially the condensate is <strong>in</strong> the ground state. In thevariational description, this translates <strong>in</strong>to <strong>in</strong>itial conditions u ρ (0) = u ρ0 , ˙u ρ (0) = 0,u z (0) = u z0 , ˙u z (0) = 0, where u ρ0 and u z0 are the time-<strong>in</strong>dependent solutions115


5. BEC excitation by modulation of scatter<strong>in</strong>g lengthAxial condensate widthRadial condensate width35302520151.61.51.41.31.21.11variationalGP numerics0 200 400 600 800 1000 1200variationalGP numerics1.21.1t1220 225 230 235 2400 50 100 150 200 250Figure 5.5: Time-dependent axial and radial condensate widths calculated as rootmean square averages. Comparison of the numerical solution of the time-dependentGP equation with a solution obta<strong>in</strong>ed by us<strong>in</strong>g the Gaussian approximation for theactual experimental parameters <strong>in</strong> Eq. (5.9) and Ω = 0.05.tof Eqs. (5.11) and (5.12), while <strong>in</strong> GP simulations we reach the ground state byperform<strong>in</strong>g an imag<strong>in</strong>ary-time propagation until convergence to the ground stateis achieved [87]. For solv<strong>in</strong>g the GP equation (4.15), we use the split-step Crank-Nicolson method [87] expla<strong>in</strong>ed <strong>in</strong> detail <strong>in</strong> Appendix A. From the typical resultspresented <strong>in</strong> Fig. 5.5, which correspond to the actual experimental parameters, itis evident that we have a good qualitative agreement between the two approaches,even for long times of the dynamical evolution.116


5. BEC excitation by modulation of scatter<strong>in</strong>g length5.4 Spherically-symmetric BECUs<strong>in</strong>g a simple symmetry-based reason<strong>in</strong>g, we immediately conclude that harmonicmodulation of the <strong>in</strong>teraction strength <strong>in</strong> the case of a spherically symmetric BEC,(λ z = 1), leads to the excitation of the breath<strong>in</strong>g mode only, so that u ρ (t) = u z (t) ≡u(t). This fact simplifies numerical and analytical calculations and this is why wefirst consider this case before we embark to the study of a more complex, axiallysymmetricBEC.The system of ord<strong>in</strong>ary differential Eqs. (5.11) and (5.12) <strong>in</strong> this case reduces toa s<strong>in</strong>gle equation:ü(t) + u(t) − 1u(t) 3 − P(t)u(t) 4 = 0 . (5.15)The equilibrium condensate width u 0 satisfies the equationu 0 − 1 u 3 0− P u 4 0= 0 , (5.16)and a l<strong>in</strong>ear stability analysis yields the breath<strong>in</strong>g mode frequencyω 0 =√1 + 3 u 4 0+ 4Pu 5 0. (5.17)Note that the above result for the breath<strong>in</strong>g mode can be also obta<strong>in</strong>ed from Eq. (5.5)if we set λ z = 1, u ρ0 = u z0 ≡ u 0 , and take <strong>in</strong>to account Eq. (5.16).The ma<strong>in</strong> feature of the modulation-<strong>in</strong>duced dynamics is that it strongly dependson the value of the driv<strong>in</strong>g frequency Ω. To illustrate this, we set P = 0.4, Q = 0.1and solve Eq. (5.15) for different values of the driv<strong>in</strong>g Ω. From the l<strong>in</strong>ear responsetheory, we have u 0 = 1.08183, ω 0 = 2.06638 and we assume that the condensate is<strong>in</strong>itially <strong>in</strong> equilibrium, i.e. u(0) = u 0 , ˙u(0) = 0. Numerical results are presented <strong>in</strong>Fig. 5.6. Large amplitude oscillations and beat<strong>in</strong>g phenomena are observed for bothΩ ≈ ω 0 and for Ω ≈ 2ω 0 .The phenomenology based on Eq. (5.15) is more systematically shown <strong>in</strong> Fig. 5.7,where we plot the oscillation amplitude, def<strong>in</strong>ed as (u max −u m<strong>in</strong> )/2, versus the driv<strong>in</strong>gfrequency Ω. A resonant behavior becomes apparent for both Ω ≈ ω 0 and Ω ≈ 2ω 0 .In the same figure we also show the expected positions of resonances calculatedus<strong>in</strong>g the l<strong>in</strong>ear stability analysis. Clearly, the prom<strong>in</strong>ent peaks exhibit shifts withrespect to the solid vertical l<strong>in</strong>es, represent<strong>in</strong>g ω 0 and 2ω 0 . As expected, a stronger117


5. BEC excitation by modulation of scatter<strong>in</strong>g lengthu(t)u(t)u(t)1.141.121.11.081.061.04Ω = 1, analyticsΩ = 1, numerics1.020 5 10 15 20 25 30 35 4065432100 50 100 150 200 250 300 350 4004.53.5 42.5 31.5 2Ω = 2.04, numericsΩ = 4.1, numericst0.5 100 50 100 150 200 250 300 350tt21.81.61.41.210.80.61.111.11.091.081.071.091.08Ω = 2, analyticsΩ = 2, numerics0 5 10 15 20 25 30 35 40Ω = 4, analyticsΩ = 4, numericst0 5 10 15 20Ω = 5, analyticsΩ = 5, numericst0 5 10 15 20tFigure 5.6: Condensate width dynamics u(t) versus t with<strong>in</strong> the Gaussian approximationfor P = 0.4, Q = 0.1 and several different driv<strong>in</strong>g frequencies Ω. We plotthe exact numerical solution of Eq. (5.15). For off-resonant driv<strong>in</strong>g frequencies Ω,we also show our analytical third-order perturbative result (designated ’analytics’),as expla<strong>in</strong>ed <strong>in</strong> section III.B.modulation amplitude leads to a larger frequency shift, as can be seen from the<strong>in</strong>set.The curves presented <strong>in</strong> Fig. 5.7 are obta<strong>in</strong>ed by an equidistant sampl<strong>in</strong>g of theexternal driv<strong>in</strong>g frequency Ω. In addition to the expected resonances close to ω 0and 2ω 0 , a more thorough exploration of solutions of the variational equation (5.15)shows that other “resonances” are present, such as, e.g. Ω ≈ ω 0 /2 and Ω ≈ 2ω 0 /3.This is further demonstrated <strong>in</strong> Fig. 5.8. These “resonances” are harder to observe118


5. BEC excitation by modulation of scatter<strong>in</strong>g length1000101001Amplitude1010.10.11.9622.042.080.01Q = 0.3Q = 0.050.0011.5 2 2.5 3 3.5 4 4.5 5Figure 5.7: Oscillation amplitude (u max − u m<strong>in</strong> )/2 versus driv<strong>in</strong>g frequency Ω forP = 0.4. In the <strong>in</strong>set, we zoom to the first peak to emphasize that both the shapeand the value of a resonance depend on the modulation amplitude Q and thatresonances occur at a driv<strong>in</strong>g frequency Ω, which differs from ω 0 . The solid verticall<strong>in</strong>es correspond to ω 0 and 2ω 0 .Ωu(t)2.521.51Ω = 1.032u(t)1.81.61.41.21Ω = 1.3750.50.80 100 200 300 400 5000 1000 2000 3000ttFigure 5.8: Exact numerical solution of Eq. (5.15) for the condensate width u(t)versus t for P = 0.4, Q = 0.3, correspond<strong>in</strong>g to ω 0 = 2.06638. We observe largeamplitude oscillations for Ω ≈ ω 0 /2 <strong>in</strong> the left panel, while <strong>in</strong> the right panel the“resonant” behavior is present for Ω ≈ 2ω 0 /3.numerically, s<strong>in</strong>ce it is necessary to perform a f<strong>in</strong>e tun<strong>in</strong>g of the external frequency.However, they clearly demonstrate nonl<strong>in</strong>ear BEC properties and an experimentalobservation of these phenomena is certa<strong>in</strong>ly of high <strong>in</strong>terest. We note that theobserved resonance pattern of the form Ω ≈ 2ω 0 /n (where n is an <strong>in</strong>teger) arisesalso <strong>in</strong> the case of a parametrically driven system described by the Mathieu equation,119


5. BEC excitation by modulation of scatter<strong>in</strong>g length10 3-40 -30 -20 -10 0 10 20 30 4010 210 110 010 -110 -210 -310 -410 -510 -6Ω = 2Frequency10 210 110 0ω - Ω10 -110 -20.05 0.06 0.07 0.08Frequency10 2 Ω + ω10 0 2Ω 2ω10 -210 -44 4.05 4.1 4.15 4.2Frequency10 210 1 Ω ω10 010 -110 -21.9 2 2.1 2.2Frequency10 110 0 2Ω + ω Ω + 2ω10 -110 -2 3Ω3ω10 -36 6.05 6.1 6.15 6.2FrequencyFigure 5.9: Fourier transform of u(t) for P = 0.4, Q = 0.1, and Ω = 2. First plotpresents the complete spectrum on a semi-log scale, while the subsequent plots focuson regions of <strong>in</strong>terest <strong>in</strong> the spectrum.for <strong>in</strong>stance, <strong>in</strong> the context of the Paul trap [130].To exam<strong>in</strong>e such excited modes directly, we look at the Fourier transform of thecondensate width u(t). To this end, we numerically solve Eq. (5.15) and f<strong>in</strong>d theFourier transform of its solution us<strong>in</strong>g the MATHEMATICA software package [54].An example of such an excitation spectrum for P = 0.4, Q = 0.1, and Ω = 2 isgiven <strong>in</strong> Fig. 5.9. The spectrum conta<strong>in</strong>s two prom<strong>in</strong>ent modes: a breath<strong>in</strong>g modeof frequency ω (close, but not equal to ω 0 ), and a mode that corresponds to thedriv<strong>in</strong>g frequency Ω, along with many higher-order harmonics of the general formmΩ + nω, where m and n are <strong>in</strong>tegers.In Fig. 5.10 we juxtapose two zoomed-<strong>in</strong> Fourier spectra for two different driv<strong>in</strong>gfrequencies for P = 0.4 and Q = 0.2. On the left plot, we show a spectrum forΩ = 1. The vertical solid l<strong>in</strong>e corresponds to ω 0 and we f<strong>in</strong>d the peak <strong>in</strong> the120


5. BEC excitation by modulation of scatter<strong>in</strong>g length10 1 1.98 2.02 2.06 2.110 010 -110 -2Ω = 1ω 010 210 110 010 -1ω 010 -2 Ω = 21.98 2.02 2.06 2.1 2.14FrequencyFrequencyFigure 5.10: Parts of the Fourier spectra for P = 0.4, Q = 0.2, and two differentdriv<strong>in</strong>g frequencies: Ω = 1 (left) and Ω = 2 (right). Position of a l<strong>in</strong>ear responseresult ω 0 is given by a vertical solid l<strong>in</strong>e.spectrum that lies almost precisely at this position. On the contrary, from the rightplot of Fig. 5.10, which corresponds almost to the resonant excitation at Ω = 2, wesee that the prom<strong>in</strong>ent peak is displaced from the vertical l<strong>in</strong>e. This is the mostclear-cut illustration of the shifted eigenfrequency aris<strong>in</strong>g due to the nonl<strong>in</strong>earityof the underly<strong>in</strong>g dynamical equations. Our objective is to develop an analyticalapproach capable of tak<strong>in</strong>g <strong>in</strong>to account these nonl<strong>in</strong>ear effects.5.4.1 Po<strong>in</strong>caré-L<strong>in</strong>dstedt methodIn its essence, our analytical approach represents the standard Po<strong>in</strong>caré-L<strong>in</strong>dstedtmethod [127, 128, 124, 126]. L<strong>in</strong>eariz<strong>in</strong>g the variational equation (5.15) around thetime-<strong>in</strong>dependent solution u 0 for a vanish<strong>in</strong>g driv<strong>in</strong>g Q = 0, we obta<strong>in</strong> the zerothorderapproximation for the collective mode ω = ω 0 , expressed by Eq. (5.17). Tocalculate the collective mode to higher orders, we explicitly <strong>in</strong>troduce the soughtaftereigenfrequency ω <strong>in</strong>to the calculation by rescal<strong>in</strong>g the time from t to s = ωt,yield<strong>in</strong>g the equation:ω 2 ü(s) + u(s) − 1u(s) 3 −Pu(s) 4 −Q Ωscosu(s)4ω = 0 . (5.18)In the next step, we assume the follow<strong>in</strong>g perturbative expansions <strong>in</strong> the modulationamplitude Q:u(s) = u 0 + Q u 1 (s) + Q 2 u 2 (s) + Q 3 u 3 (s) + . . . , (5.19)ω = ω 0 + Q ω 1 + Q 2 ω 2 + Q 3 ω 3 + . . . , (5.20)121


5. BEC excitation by modulation of scatter<strong>in</strong>g lengthwhere we expand ω around ω 0 and <strong>in</strong>troduce the frequency shifts ω 1 , ω 2 , . . ., for eachorder <strong>in</strong> the expansion <strong>in</strong> Q. By <strong>in</strong>sert<strong>in</strong>g the above expansions <strong>in</strong>to the Eq. (5.18)and collect<strong>in</strong>g terms of the same order <strong>in</strong> Q, we obta<strong>in</strong> a hierarchical system of l<strong>in</strong>eardifferential equations. To the third order, we f<strong>in</strong>d:ω 2 0ü1(s) + ω 2 0 u 1(s) = 1 u 4 0cos Ωsω ,ω0ü2(s) 2 + ω0 2 u 2(s) = −2ω 0 ω 1 ü 1 (s) − 4 uu 5 1 (s) cos Ωs0 ω + αu 1(s) 2 ,ω0ü3(s) 2 + ω0 2 u 3(s) = −2ω 0 ω 2 ü 1 (s) − 2βu 1 (s) 3 + 2αu 1 (s)u 2 (s) − ω1ü1(s)2+ 10 uu 6 1 (s) 2 cos Ωs0 ω − 4 uu 5 2 (s) cos Ωs0 ω − 2ω 0 ω 1 ü 2 (s),where α = 10 P/u 6 0 + 6/u5 0 and β = 10 P/u7 0 + 5/u6 0 .These equations disentangle <strong>in</strong> a natural way: we solve the first one for u 1 (s), anduse that solution to solve the second one for u 2 (s), and so on. At the n-th level of theperturbative expansion (n ≥ 1) we use the <strong>in</strong>itial conditions u n (0) = 0, ˙u n (0) = 0.As is well known, the presence of the term cos s on the right-hand side of some ofthe previous equations would yield a solution that conta<strong>in</strong>s the secular term s s<strong>in</strong> s.Such a secular term grows l<strong>in</strong>early <strong>in</strong> time, which makes it the dom<strong>in</strong>ant term <strong>in</strong>the expansion (5.19) that otherwise conta<strong>in</strong>s only periodic functions <strong>in</strong> s. In orderto ensure a regular behavior of the perturbative expansion, the respective frequencyshifts ω 1 , ω 2 , . . . are determ<strong>in</strong>ed by impos<strong>in</strong>g the cancellation of secular terms.This analytical procedure is implemented up to the third order <strong>in</strong> the modulationamplitude Q by us<strong>in</strong>g the software package MATHEMATICA [54]. Although thecalculation is straightforward, it easily becomes tedious for higher orders of perturbationtheory. Note that it is necessary to perform the calculation to at least thirdorder s<strong>in</strong>ce it turns out to be the lowest-order solution where secular terms appearand where the nontrivial frequency shift can be calculated. We have solved explicitlyequations for u 1 (s), u 2 (s), and u 3 (s) and <strong>in</strong> Fig. 5.6 we show an excellent agreementof our analytical solutions with a respective numerical solution of Eq. (5.15). Fromthe first-order solution u 1 (t) we read off only the two basic modes ω 0 and Ω, whilethe second-order harmonics 2ω 0 , ω 0 − Ω, ω 0 + Ω and 2Ω appear <strong>in</strong> u 2 (t). In thethird order of perturbation theory, higher-order harmonics ω − 2Ω, 2ω − Ω, 2ω + Ω,ω + 2Ω, 3ω, and 3Ω are also present. Concern<strong>in</strong>g the cancellation of secular terms,122


5. BEC excitation by modulation of scatter<strong>in</strong>g lengththe first-order correction ω 1 vanishes, lead<strong>in</strong>g to a frequency shift which is quadratic<strong>in</strong> Q:ω = ω 0 +Q212u 200 ω3 0where the polynomial P(Ω) is given byP(Ω)(Ω 2 − ω 2 0 )2 (Ω 2 − 4ω 2 0 ) + . . . , (5.21)P(Ω)=Ω 4 [ −240Pu 5 0 + 36u6 0 (−4 + 3u4 0 ω2 0 )]+Ω 2 [ −1100P 2 − 30Pu 0 (44 − 65u 4 0 ω2 0 ) + 9u2 0 (−44 + 127u4 0 ω2 0 − 44u8 0 ω4 0 )]+5600P 2 ω 2 0 + 840Pu 0 ω 2 0(8 − 3u 4 0ω 2 0) + 36u 2 0ω 2 0(56 − 39u 4 0ω 2 0 + 8u 8 0ω 4 0). (5.22)MATHEMATICA notebook which implements this analytical calculation is availableat our web site [60].5.4.2 Results and discussionThe result given by Eq. (5.21) is the ma<strong>in</strong> achievement of our analytical analysis <strong>in</strong>the previous section. It is obta<strong>in</strong>ed with<strong>in</strong> a second-order perturbative approach <strong>in</strong>Q and it describes the breath<strong>in</strong>g mode frequency dependence on the driv<strong>in</strong>g Ω andthe modulation amplitude Q as a result of nonl<strong>in</strong>ear effects. Due to the underly<strong>in</strong>gperturbative expansion, we do not expect Eq. (5.21) to be mean<strong>in</strong>gful at the preciseposition of the resonances. However, by comparison with numerical results basedon the variational equation, we f<strong>in</strong>d that Eq. (5.21) represents quite a reasonableapproximation even close to the resonant region.To illustrate this, <strong>in</strong> Fig. 5.11 we show two such comparisons. In the upperpanel we consider the parameter set P = 0.4 and Q = 0.1, and observe significantfrequency shifts only <strong>in</strong> the narrow resonant regions. We notice an excellent agreementof numerical values with the analytical result given by Eq. (5.21). In the lowerpanel we consider the parameter set P = 1 and Q = 0.8, with much stronger modulationamplitude. In this case we observe significant frequency shifts for the broaderrange of modulation frequencies Ω. In spite of a strong modulation, we still seea qualitatively good agreement of numerical results with the analytical predictiongiven by Eq. (5.21). In pr<strong>in</strong>ciple, better agreement can be achieved us<strong>in</strong>g higherorderperturbative approximation. The dashed l<strong>in</strong>e on both figures represents Ω/2,given as a guide to the eye. It also serves as a crude description of what we observenumerically <strong>in</strong> the range Ω ≈ 2ω 0 .123


5. BEC excitation by modulation of scatter<strong>in</strong>g length2.12.08ω 0analytical ωnumerical ωFrequency2.062.042.02Frequency22.32.252.22.152.12.0521.951.91 1.5 2 2.5 3 3.5 4 4.5 5ω 0analytical ωnumerical ωΩ1 2 3 4 5 6ΩFigure 5.11: Frequency of the breath<strong>in</strong>g mode versus the driv<strong>in</strong>g frequency Ω forP = 0.4 and Q = 0.1 (top), and P = 1 and Q = 0.8 (bottom). The dashed l<strong>in</strong>erepresents Ω/2 and is given to guide the eye.The presence of two poles at Ω = ω 0 and Ω = 2ω 0 <strong>in</strong> Eq. (5.21) implies thepossible existence of real resonances <strong>in</strong> the BEC with a harmonically modulated<strong>in</strong>teraction. A perturbative expansion to higher orders would <strong>in</strong>troduce some additionalpoles, responsible for higher-order “resonant” behavior observed at Ω ≈ 2ω 0 /n(n ≥ 3). Still, the poles seem to be only an artifact of our approximative perturbativescheme, not present <strong>in</strong> the exact description. For example, a simple resummationperformed us<strong>in</strong>g the second-order perturbative result removes these effects, although124


5. BEC excitation by modulation of scatter<strong>in</strong>g lengththis is only an ad-hoc approximation. We stress that this issue concern<strong>in</strong>g the trueresonant behavior can not be settled either by rely<strong>in</strong>g on a numerical calculationdue to <strong>in</strong>herent numerical artifacts related to f<strong>in</strong>ite numerical precision and f<strong>in</strong>itecomputational time. To resolve it, one should rely on an analytical considerationalong the l<strong>in</strong>es of Ref. [118] or use some analytical tool applicable at resonances,such as the resonant Bogoliubov-Mitropolsky method [124].10 3 1.94 1.96 1.98 2 2.02 2.04 2.06 2.0810 210 110 010 -110 -210 -3GP numerics 1GP numerics 2GP numerics 3Gaussian app.ΩFrequencyωFigure 5.12: Part of the Fourier spectrum of the time-dependent condensate widthfor P = 0.4, Q = 0.2, Ω = 2. For numerical solution of GP equation we use severaldiscretization schemes: GP numerics 1 (time step ε = 10 −3 , spac<strong>in</strong>g h = 4 × 10 −2 ),GP numerics 2 (ε = 5 × 10 −4 , h = 2 × 10 −2 ), GP numerics 3 (ε = 5 × 10 −5 ,h = 5 × 10 −3 ). Details of the used numerical algorithms are given <strong>in</strong> AppendixA. For comparison we also show the correspond<strong>in</strong>g spectrum obta<strong>in</strong>ed from theGaussian approximation (dotted-dashed l<strong>in</strong>e) and analytical result (5.21) for theposition of breath<strong>in</strong>g mode (solid vertical l<strong>in</strong>e).In addition to comparison of our analytical results with numerical solutions basedon the Gaussian variational approximation, we present a comparison with the fullnumerical solution of the GP equation. In order to be able to perform Fourieranalysis with sufficient resolution, it is necessary to obta<strong>in</strong> an accurate solutionfor long evolution times. We do this by us<strong>in</strong>g the split-step method <strong>in</strong> comb<strong>in</strong>ationwith the semi-implicit Crank-Nicolson method [87]. As we ref<strong>in</strong>e the GP numerics byus<strong>in</strong>g f<strong>in</strong>er space and time discretization parameters, our numerical results becomestable as shown <strong>in</strong> Fig. 5.12. From the same figure, we observe quantitatively goodagreement between GP numerics and Gaussian approximation, reflected <strong>in</strong> close125


5. BEC excitation by modulation of scatter<strong>in</strong>g lengthvalues obta<strong>in</strong>ed for the breath<strong>in</strong>g mode frequency. In addition, numerical values forthe breath<strong>in</strong>g mode approach closely the analytical result of Eq. (5.21), shown by asolid vertical l<strong>in</strong>e <strong>in</strong> Fig. 5.12.It is well known that for a correspond<strong>in</strong>g two-dimensional axially-symmetricsystem with a constant <strong>in</strong>teraction and trapp<strong>in</strong>g frequency, the breath<strong>in</strong>g modeoscillations can be described by an exact l<strong>in</strong>ear equation [131, 132]. However, <strong>in</strong>the case of a time-dependent trapp<strong>in</strong>g frequency, the exact equation of motion isnonl<strong>in</strong>ear [118]. To the best of our knowledge, for a time-dependent <strong>in</strong>teractionstrength the correspond<strong>in</strong>g exact equation does not exist <strong>in</strong> the literature, but onecan reasonably expect that nonl<strong>in</strong>ear effects will rema<strong>in</strong> <strong>in</strong> such systems, due to the<strong>in</strong>herent time dependence of the <strong>in</strong>teraction.5.5 Axially-symmetric BECTo obta<strong>in</strong> results relevant for a comparison with the experiment reported <strong>in</strong> Ref. [15],we now study an axially-symmetric BEC. An illustration of the condensate dynamicsis shown <strong>in</strong> Fig. 5.13 for P = 1, Q = 0.2, λ z = 0.3. We plot numerical solutions ofEqs. (5.11) and (5.12) obta<strong>in</strong>ed for the equilibrium <strong>in</strong>itial conditions u ρ (0) = u ρ0 ,˙u ρ (0) = 0, u z (0) = u z0 , and ˙u z (0) = 0. For the specified parameters, the equilibriumwidths are found to be u ρ0 = 1.09073, u z0 = 2.40754, and from the l<strong>in</strong>ear stabilityanalysis we f<strong>in</strong>d both the quadrupole mode frequency ω Q0 = 0.538735 and thebreath<strong>in</strong>g mode frequency ω B0 = 2.00238. For a driv<strong>in</strong>g frequency Ω close to ω Q0 , weobserve large amplitude oscillations <strong>in</strong> the axial direction. An example of excitationspectra is shown <strong>in</strong> Fig. 5.14. Here, we have the three basic modes ω Q , ω B , Ω, andmany higher-order harmonics.5.5.1 Po<strong>in</strong>caré-L<strong>in</strong>dstedt methodIn order to extract <strong>in</strong>formation on the frequencies of the collective modes beyondthe l<strong>in</strong>ear stability analysis, we apply the perturbative expansion <strong>in</strong> the modulationamplitude Q:u ρ (t) = u ρ0 + Q u ρ1 (t) + Q 2 u ρ2 (t) + Q 3 u ρ3 (t) + . . ., (5.23)u z (t) = u z0 + Q u z1 (t) + Q 2 u z2 (t) + Q 3 u z3 (t) + . . .. (5.24)126


5. BEC excitation by modulation of scatter<strong>in</strong>g lengthu z (t)32.82.62.42.2Ω = 0.4, analyticsΩ = 0.4, numerics20 10 20 30 40 50 60 70 80tu z (t)2.552.52.452.42.35Ω = 1, analyticsΩ = 1, numerics0 10 20 30 40 50 60 70 80tu ρ (t)1.151.141.131.121.111.11.091.081.071.061.05Ω = 0.4, analyticsΩ = 0.4, numerics0 10 20 30 40 50 60 70 80tu ρ (t)1.161.141.121.11.081.061.04Ω = 1, analyticsΩ = 1, numerics0 5 10 15 20 25 30 35 40tFigure 5.13: Condensate dynamics with<strong>in</strong> the Gaussian approximation for P = 1,Q = 0.2, λ z = 0.3 and two different driv<strong>in</strong>g frequencies Ω = 0.4 (left plot) and Ω = 1(right plot). We plot exact numerical solution of Eqs. (5.11) and (5.12) together withthe analytical second-order perturbative result, as expla<strong>in</strong>ed <strong>in</strong> section IV.A.10 310 2 FT of u zFT of u10 1ρ 10 110 010 010 -110 -210 -110 -310 -210 -4-4 -2 0 2 410 -3ΩFT of u zω Qω B10 2 0.5 1 1.5 2FrequencyFrequencyFigure 5.14: Fourier transformed u ρ (t) and u z (t) for P = 1, Q = 0.2, λ z = 0.3, andΩ = 0.4. Left plot gives complete spectrum, while on the right plot we show part ofthe spectrum together with positions of prom<strong>in</strong>ent peaks.By <strong>in</strong>sert<strong>in</strong>g these expansions <strong>in</strong> Eqs. (5.11) and (5.12), and by perform<strong>in</strong>g additionalexpansions <strong>in</strong> Q, we obta<strong>in</strong> a system of l<strong>in</strong>ear differential equations of the generalform:ü ρn (t) + m 11 u ρn (t) + m 12 u zn (t) + f ρn (t) = 0, (5.25)m 21 u ρn (t) + ü zn (t) + m 22 u zn (t) + f zn (t) = 0, (5.26)127


where n = 1, 2, 3, . . . is an <strong>in</strong>teger, and5. BEC excitation by modulation of scatter<strong>in</strong>g lengthm 11 = 4, m 12 = P/u 3 ρ0 u2 z0 , m 21 = 2 P/u 3 ρ0 u2 z0 , m 22 = λ 2 z + 3/u4 z0 + 2 P/u2 ρ0 u3 z0 .The n-th order functions f ρn (t) and f zn (t) depend only on the solutions u ρi (t) andu zi (t) of lower orders, i < n. For n = 1 we haveand for n = 2 we get correspond<strong>in</strong>glycos Ωt cos Ωtf ρ1 (t) = −u 3 ρ0 u , f z1 (t) = − ,z0u 2 ρ0 u2 z0f ρ2 (t) =f z2 (t) =3cos Ωt uu 4 ρ1 (t) − 6 uρ0u z0 u 5 ρ1 (t) 2 −6P uρ0 u 5 ρ1 (t) 2ρ0u z0+ 1 cos Ωt uu 3 ρ0u 2 z1 (t) −3P uz0u 4 ρ0u 2 ρ1 (t)u z1 (t) −P uz0 u 3 ρ0u 3 z1 (t) 2 ,z02cos Ωt u ρ1 (t) −3P u ρ1 (t) 2 −3P u z1 (t) 2u 3 ρ0 u2 z0u 4 ρ0 u2 z0+ 2 cos Ωt uu 2 z1 (t) − 6 uρ0 u3 z0u 5 z1 (t) 2 −z0u 2 ρ0 u4 z04Pu 3 ρ0 u3 z0u ρ1 (t)u z1 (t).The l<strong>in</strong>ear transformationu ρn (t) = x n (t) + y n (t), (5.27)u zn (t) = c 1 x n (t) + c 2 y n (t), (5.28)with the coefficientsc 1,2 = m 22 − m 11 ± √ (m 22 − m 11 ) 2 + 4m 12 m 212m 12,decouples the system at the n-th level and leads to the equations of the form:ẍ n (t) + ω 2 Q0 x n(t) + c 2 f ρn (t) − f zn (t)c 2 − c 1= 0, (5.29)ÿ n (t) + ω 2 B0 y n(t) + c 1 f ρn (t) − f zn (t)c 1 − c 2= 0. (5.30)Now it is clear how to proceed: we first solve Eqs. (5.29) and (5.30) for x 1 (t) andy 1 (t), and then us<strong>in</strong>g Eqs. (5.27) and (5.28) we obta<strong>in</strong> u ρ1 (t) and u z1 (t). In the128


5. BEC excitation by modulation of scatter<strong>in</strong>g lengthnext step, we use these solutions and solve for x 2 (t) and y 2 (t) and so on. At eachlevel n ≥ 1 we impose the <strong>in</strong>itial conditions u ρn (0) = 0, ˙u ρn (0) = 0, u zn (0) = 0,and ˙u zn (0) = 0. At the first level of perturbation theory, equations for x and yare decoupled, i.e. x 1 (t) and y 1 (t) are normal modes: x 1 (t) describes quadrupoleoscillations, while y 1 (t) describes breath<strong>in</strong>g oscillations. However, at the secondorder of perturbation theory y 1 (t) enters the equation for x 2 (t) and also x 1 (t) appears<strong>in</strong> equation for y 2 (t), i.e. we have a nonl<strong>in</strong>ear mode coupl<strong>in</strong>g.We have performed the explicit calculation to the second order by us<strong>in</strong>g thesoftware package MATHEMATICA [54]. We have obta<strong>in</strong>ed an excellent agreementof second-order analytical results and numerical results, as can be seen <strong>in</strong> Fig. 5.13for a moderate value of a modulation amplitude Q. The first secular terms appearat the level n = 3. The expressions are cumbersome, but the relevant behavior isobta<strong>in</strong>ed from the follow<strong>in</strong>g terms <strong>in</strong> the equation for x 3 (t):ẍ 3 (t) + ωQ0x 2 3 (t) + C Q cos ω Q0 t + . . . = 0, (5.31)which leads tox 3 (t) = − C Q2ω Q0t s<strong>in</strong> ω Q0 t + . . . (5.32)The last term can be absorbed <strong>in</strong>to the first-order solutionu ρ (t) = A Q cos ω Q0 t − C QQ 2t s<strong>in</strong> ω Q0 t + . . .2ω Q0≈ A Q cos [(ω Q0 + ∆ω Q0 )t], (5.33)and can be <strong>in</strong>terpreted as a frequency shift of the quadrupole mode, quadratic <strong>in</strong> Q:ω Q = ω Q0 + ∆ω Q0 = ω Q0 + Q 2 + . . . (5.34)2ω Q0 A QThe coefficients A Q and C Q are calculated us<strong>in</strong>g the MATHEMATICA code availableat our site [60], and their explicit form is too long to be presented here. Along thesame l<strong>in</strong>es we have also calculated the frequency shift of the breath<strong>in</strong>g mode.C Q5.5.2 Results and discussionThe ma<strong>in</strong> results of our calculation <strong>in</strong> this section are shown <strong>in</strong> Figs. 5.15 and5.16. In Fig. 5.15 we plot the analytically obta<strong>in</strong>ed frequency of the quadrupole129


5. BEC excitation by modulation of scatter<strong>in</strong>g length0.55Frequency ω Q0.5450.540.5350.530 1 2 3 4 5Ωω Q0ω Q, (a)ω Q, (n)Figure 5.15: Frequency of the quadrupole mode ω Q versus driv<strong>in</strong>g frequency Ω forP = 1, Q = 0.2, and λ z = 0.3. We plot l<strong>in</strong>ear response result ω Q0 , second-orderanalytical result ω Q,(a) and numerical values ω Q,(n) .Frequency ω B2.012.0082.0062.0042.00221.9981.9961.9941.9921.990 1 2 3 4 5 6Ωω B0ω B, (a)ω B, (n)Figure 5.16: Frequency of the breath<strong>in</strong>g mode ω B versus driv<strong>in</strong>g frequency Ω forP = 1, Q = 0.2, and λ z = 0.3. We plot l<strong>in</strong>ear response result ω B0 , second-orderanalytical result ω B,(a) and numerical values ω B,(n) .mode versus the driv<strong>in</strong>g Ω, us<strong>in</strong>g the second order perturbation theory togetherwith the correspond<strong>in</strong>g numerical result based on the Fourier analysis of solutionsof Eqs. (5.11) and (5.12). An analogous plot for the frequency of the breath<strong>in</strong>g mode130


5. BEC excitation by modulation of scatter<strong>in</strong>g lengthis given <strong>in</strong> Fig. 5.16. Our analytical perturbative result for the shifted quadrupolemode frequency conta<strong>in</strong>s poles at ω Q0 , 2ω Q0 , ω B0 −ω Q0 , ω Q0 +ω B0 and ω B0 . Similarly,for the shifted frequency of the breath<strong>in</strong>g mode poles <strong>in</strong> the perturbative solutionare found at ω Q0 , ω B0 , 2ω B0 , ω B0 − ω Q0 and ω Q0 + ω B0 . In both figures we seeexcellent agreement of the perturbatively obta<strong>in</strong>ed results with the exact numerics.In the experiment from Ref. [15], excitations of a highly elongated and stronglyrepulsive BEC were considered with the system parameters given <strong>in</strong> Eq. (5.9). Forthat case, accord<strong>in</strong>g to Eq. (5.10) we get ω Q0 ≪ ω B0 , and the driv<strong>in</strong>g frequencywas chosen <strong>in</strong> the range (0, 3ω Q0 ). Good agreement of real-time dynamics obta<strong>in</strong>edfrom the variational approximation with the exact solution of the time-dependentGP simulation occurs even for long propagation times, as can be seen <strong>in</strong> Fig. 5.5,which implies a good accuracy of the Gaussian approximation for calculat<strong>in</strong>g thefrequencies of the excited modes. From the real-time dynamics shown <strong>in</strong> Fig. 5.5, weobserve the excitation of the slow quadrupole mode as an out-of phase oscillation <strong>in</strong>the axial and <strong>in</strong> the radial direction. In addition, <strong>in</strong> the radial direction we observefast breath<strong>in</strong>g mode oscillations. This is typical for highly elongated condensates[133] and our analysis for the experimental parameters shows a strong excitationof the quadrupole mode, but also a significant excitation of the breath<strong>in</strong>g mode<strong>in</strong> the radial direction. Due to the large modulation amplitude Q, many higherorder harmonics are excited, and, most importantly, we f<strong>in</strong>d frequency shifts of thequadrupole mode of about 10% <strong>in</strong> Fig 5.17. From the same figure we notice that,due to the chosen frequency range for Ω, only resonances located at ω Q and 2ω Qare observed. The presence of nonl<strong>in</strong>ear effects is already mentioned <strong>in</strong> Ref. [15].However, we conclude that frequency shifts calculated here have to be taken <strong>in</strong>toaccount for extract<strong>in</strong>g the resonance curves from the underly<strong>in</strong>g experimental data.To achieve more clear-cut experimental observation of the nonl<strong>in</strong>earity-<strong>in</strong>ducedfrequency shifts calculated <strong>in</strong> this paper, we suggest a different trap geometry fromthe one used <strong>in</strong> Ref. [15]. Measurements of stable BEC modes can be performedfor about 1 s, and <strong>in</strong> order to extract precise values of the excited frequencies <strong>in</strong>the Fourier analysis, several oscillation periods should be captured with<strong>in</strong> this time<strong>in</strong>terval. A higher frequency of the quadrupole mode, that can be realized by us<strong>in</strong>ga larger trap aspect ratio λ z , <strong>in</strong> comb<strong>in</strong>ation with a higher modulation frequencywould fulfill this condition. Accord<strong>in</strong>g to the results presented <strong>in</strong> Ref. [15], resonantdriv<strong>in</strong>g may lead to condensate fragmentation. However, our numerical results<strong>in</strong>dicate frequency shifts of 10 % even outside the resonant regions accord<strong>in</strong>g to131


5. BEC excitation by modulation of scatter<strong>in</strong>g length0.040.038Frequency ω Q0.0360.0340.0320.03ω Q0ω Q, (a)ω Q, (n)0 0.02 0.04 0.06 0.08 0.1 0.12ΩFigure 5.17: Frequency of the quadrupole mode ω Q versus driv<strong>in</strong>g frequency Ω forthe experimental parameters from Eq. (5.9). We plot l<strong>in</strong>ear response result ω Q0 ,second-order analytical result ω Q,(a) , and numerical values ω Q,(n) .Figs. 5.11 and 5.17, and this is where experimental measurements should be performed.Although an <strong>in</strong>crease <strong>in</strong> λ z leads to a more pronounced nonl<strong>in</strong>ear mix<strong>in</strong>g ofquadrupole and breath<strong>in</strong>g mode and may complicate condensate dynamics further,it may be possible to perform a Fourier analysis of experimental data, analogous toRef. [101], and to compare it with the excitation spectra presented here. To achievea complete match<strong>in</strong>g of experimental data and our calculations, it may turn out thathigher-order corrections to Eq. (5.7), which arise due to the nonl<strong>in</strong>ear dependenceof scatter<strong>in</strong>g length on the external magnetic field, have to be taken <strong>in</strong>to account.5.6 Conclusions and outlookMotivated by recent experimental results, <strong>in</strong> this Chapter, we have studied nonl<strong>in</strong>earBEC dynamics <strong>in</strong>duced by a harmonically modulated <strong>in</strong>teraction at zero temperature.We have used a comb<strong>in</strong>ation of an analytic perturbative approach, numericalanalysis based on Gaussian approximation, and numerical simulations of a full timedependentGross-Pitaevskii equation. We have presented numerically calculated relevantexcitation spectra and found prom<strong>in</strong>ent nonl<strong>in</strong>ear features: mode coupl<strong>in</strong>g,higher harmonics generation, and significant shifts <strong>in</strong> the frequencies of collectivemodes. In addition, we have provided an analytical perturbative framework that132


5. BEC excitation by modulation of scatter<strong>in</strong>g lengthcaptures most of the observed phenomena. The ma<strong>in</strong> results are analytic formulaedescrib<strong>in</strong>g the dependence of collective mode frequencies on the modulation amplitudeand on the external driv<strong>in</strong>g frequency for different trap geometries. Webelieve that the study presented <strong>in</strong> this Chapter is a step toward understand<strong>in</strong>gresonant processes <strong>in</strong> a nonl<strong>in</strong>ear system. To extend the applicability of our analyticalapproach, a perturbative expansion to higher order has to be performed, or anappropriate resummation of the perturbative series could be applied.The presented results could contribute to future experimental designs that may<strong>in</strong>clude mixtures of cold gases and their dynamical response to harmonically modulated<strong>in</strong>teractions, such as pattern formation <strong>in</strong>duced by the modulation of differenttime dependence of the scatter<strong>in</strong>g length. In addition, our results could contributeto resolv<strong>in</strong>g beyond-mean-field effects <strong>in</strong> the collective mode frequencies, as proposed<strong>in</strong> Refs. [134, 135], and for dipolar BEC <strong>in</strong> [136]. Nonl<strong>in</strong>earity-<strong>in</strong>duced shiftsof collective modes have to be properly taken <strong>in</strong>to account to clearly del<strong>in</strong>eate themfrom beyond-mean-field effects.133


Chapter 6SummaryS<strong>in</strong>ce the first experimental observation of Bose-E<strong>in</strong>ste<strong>in</strong> condensation <strong>in</strong> thedilute vapors of alkali atoms <strong>in</strong> 1995, the field of ultracold atoms constantly expands.The experimental advances <strong>in</strong> the manipulation of quantum gases pave a way to thedevelopment of new technologies, and allow exploration of the quantum world withpreviously unseen precision and flexibility, uncover<strong>in</strong>g at the same time new varietyof far reach<strong>in</strong>g physical phenomena. In this <strong>thesis</strong> we have studied two particularphenomena that give new <strong>in</strong>sights <strong>in</strong>to the properties of cold quantum gases andhave been recently addressed experimentally.To beg<strong>in</strong> with, <strong>in</strong> Chapter 1 we have <strong>in</strong>troduced the aspects of the research <strong>in</strong> thefield of ultracold atoms, ma<strong>in</strong> concepts, and its position with<strong>in</strong> the wide forefrontof modern physics. In Chapter 2 we have developed an efficient numerical methodfor calculation of the eigenspectrum and eigenvectors of a system <strong>in</strong> an arbitrarytrapp<strong>in</strong>g potential. The devised approach is based on the exact diagonalizationof the time-evolution operator and can be applied for numerical studies of generalfew-body systems. We have carefully explored different types of systematic errorsthat arise <strong>in</strong> the process of the spatial discretization of the time-evolution operator.We have shown analytically and numerically that the discretization error vanishesas the exponential of 1/∆ 2 , where ∆ represents the discretization spac<strong>in</strong>g. Thus,the method highly outperforms <strong>in</strong> efficiency the approaches based on the real- spacediscretization of the Hamiltonian, which exhibit polynomial error <strong>in</strong> ∆. For thehighly accurate calculation of matrix elements of the evolution operator, necessaryfor the application of the method, we use the short-time expansion of transitionamplitudes <strong>in</strong> the propagation time to high-orders. The chief <strong>in</strong>gredients of this partof the procedure are higher-order effective actions, that were derived previously. Wehave demonstrated the advantages of this method by calculat<strong>in</strong>g highly accurateenergy spectra <strong>in</strong> a numerically efficient way for several one- and two-dimensionalmodels.Motivated by experimental studies of rotat<strong>in</strong>g ultra-cold quantum gases, <strong>in</strong> Chap-134


ter 3 we have studied the effects of the shape of trapp<strong>in</strong>g potentials to the propertiesof a Bose-E<strong>in</strong>ste<strong>in</strong> condensate of an ideal gas. A rotation gives rise to the appearanceof a deconf<strong>in</strong><strong>in</strong>g harmonic potential. Particularly, <strong>in</strong> the <strong>in</strong>terest<strong>in</strong>g fast-rotat<strong>in</strong>gregime, when the rotation frequency approaches the conf<strong>in</strong><strong>in</strong>g frequency of the harmonictrap, the gas becomes unconf<strong>in</strong>ed and the condensate would disperse. Oneway to mitigate the deconf<strong>in</strong><strong>in</strong>g effect is to use an additional, quartic term <strong>in</strong> thepotential. As a result, the total external potential <strong>in</strong> the co-rotat<strong>in</strong>g frame acquiresdifferent shapes, depend<strong>in</strong>g on the rotation frequency. It is important to understandhow this affects the properties of a BEC for the <strong>in</strong>terpretation of data obta<strong>in</strong>ed <strong>in</strong>experiments with fast rotat<strong>in</strong>g BECs. By employ<strong>in</strong>g the method of exact diagonalizationof the time-evolution operator from Chapter 2, we have obta<strong>in</strong>ed numericallyexact energy spectra of the harmonic plus quartic trapp<strong>in</strong>g potential for differentvalues of the rotation frequency. Us<strong>in</strong>g this, we have calculated the condensationtemperature and found that it decreases with an <strong>in</strong>crease of the rotation frequency.We have also presented density profiles of the condensate and thermal cloud at differenttemperatures and have simulated the time-of-flight imag<strong>in</strong>g procedure for thissetup. Interest<strong>in</strong>g expansion dynamics has been found for the external potential <strong>in</strong>the shape of a Mexican hat, <strong>in</strong> the over-critical rotation regime. In the <strong>in</strong>itial stage ofthe expansion the gas expands <strong>in</strong>wards, <strong>in</strong>to the previously unoccupied <strong>in</strong>ner space,and only after that the common free expansion starts. This leads to an <strong>in</strong>crease <strong>in</strong>the typical time scales for the expansion of about one order of magnitude.Chapter 4 is dedicated to the review of the mean-field description of a weakly <strong>in</strong>teract<strong>in</strong>gBEC. We have presented several widely used approximation techniques. Inthe zero temperature limit, we have <strong>in</strong>troduced nonl<strong>in</strong>ear mean-field Gross-Pitaevskiiequation. In order to study BEC at f<strong>in</strong>ite temperature and to explore the BEC phasediagram, we have used Hartree-Fock framework <strong>in</strong> the form <strong>in</strong> which higher, thermallyexcited states are treated with<strong>in</strong> the semiclassical approximation. With<strong>in</strong> thismean-field picture, a two-component model of a BEC naturally arises. In an approximativeway, the condensate and thermal component are <strong>in</strong>troduced enabl<strong>in</strong>g us tokeep the <strong>in</strong>tuition built on a non<strong>in</strong>teract<strong>in</strong>g model. Depend<strong>in</strong>g on whether the <strong>in</strong>teractionwith<strong>in</strong> each component is taken <strong>in</strong>to account and how their mutual <strong>in</strong>teractionis considered, several approximation schemes come <strong>in</strong>to play. We have comparedtheir properties, and emphasized the identified drawbacks. The <strong>in</strong>teraction-<strong>in</strong>ducedshift of the condensation temperature has been re-derived. With these results, wehave shown how the non<strong>in</strong>teract<strong>in</strong>g BEC picture is modified <strong>in</strong> the presence of weak135


short-range <strong>in</strong>teractions.The excitation of collective oscillation modes is a direct way to probe the propertiesof a system, s<strong>in</strong>ce frequencies of collective modes can be measured very accurately,with errors of below 1%. In Chapter 5 we have studied properties of collectivemodes subject to the harmonic modulation of the s-wave scatter<strong>in</strong>g length, i.e. contact<strong>in</strong>teraction strength. We have used time-dependent Gross-Pitaevskii equationand the variational approach based on a Gaussian Ansatz to describe condensatedynamics numerically. In the non-resonant regime, when the driv<strong>in</strong>g frequency doesnot match any of the eigenfrequencies of the condensate, we have found small amplitudeoscillations that correspond to the quadrupole and breath<strong>in</strong>g mode, as expectedfor an axially-symmetric condensate. As the resonant regime is approached, nonl<strong>in</strong>eardynamical features emerge <strong>in</strong> the excitation spectra: nonl<strong>in</strong>ear mode coupl<strong>in</strong>g,higher-harmonics generation and shifts <strong>in</strong> the frequencies of excited modes. Wehave developed a perturbative approach <strong>in</strong> the modulation amplitude, based on thePo<strong>in</strong>caré-L<strong>in</strong>dstedt method, and have obta<strong>in</strong>ed analytical results for the mentionedeffects.At the end of each chapter, we have <strong>in</strong>dicated a possible future research directionto further extend the study of the topics presented <strong>in</strong> this <strong>thesis</strong>.136


Appendix A Numerical solution of the GP equationFor the numerical solution of the GP equation, we use algorithms described<strong>in</strong> Ref. [87]. Orig<strong>in</strong>al codes are written us<strong>in</strong>g the Fortran programm<strong>in</strong>g language,while we implemented the numerical procedure <strong>in</strong> the C programm<strong>in</strong>g language. Forcompleteness, <strong>in</strong> this Appendix we outl<strong>in</strong>e the ma<strong>in</strong> steps of the applied numericalmethod for the simplest, spherically-symmetric case.In general, there are two different type of questions that we want to answer bysolv<strong>in</strong>g the GP equation: either we are <strong>in</strong>terested <strong>in</strong> the equilibrium configurations,i.e. stationary solutions, or we simulate real-time dynamics of the system. It turnsout that both situations can be treated on an equal foot<strong>in</strong>g by us<strong>in</strong>g propagation <strong>in</strong>real and imag<strong>in</strong>ary time, t and τ respectively, which are connected by the expression:i t = τ .(A.1)The imag<strong>in</strong>ary-time propagation is very useful and efficient technique for obta<strong>in</strong><strong>in</strong>gstationary states of both l<strong>in</strong>ear and nonl<strong>in</strong>ear systems. Essentially, it is equivalentto the m<strong>in</strong>imization of the energy functional, and here we expla<strong>in</strong> its basics for thecase of a l<strong>in</strong>ear system. The ma<strong>in</strong> underly<strong>in</strong>g identity is given by|ψ 0 〉 = limτ→∞e −τĤ|ψ <strong>in</strong>itial 〉 ,(A.2)where |ψ <strong>in</strong>itial 〉 is an arbitrary <strong>in</strong>itial state, which has a nonzero overlap with theground-state |ψ 0 〉 of a system. Eq. (A.2) states that after long enough propagation<strong>in</strong> the imag<strong>in</strong>ary time, we will obta<strong>in</strong> the ground state of the system. This canbe easily understood by decompos<strong>in</strong>g the <strong>in</strong>itial state <strong>in</strong>to the eigenvectors of theHamiltonian Ĥ, e −τĤ|ψ ∞∑<strong>in</strong>itial 〉 = 〈ψ k |ψ <strong>in</strong>itial 〉e −τE k|ψ k 〉 , (A.3)k=0by not<strong>in</strong>g that the coefficients <strong>in</strong> front of all eigenstates decay exponentially <strong>in</strong>the imag<strong>in</strong>ary time τ, and that the slowest decay<strong>in</strong>g coefficient is the one <strong>in</strong> front137


of the ground state. Imag<strong>in</strong>ary-time propagation does not preserve the norm ofthe state, and we have to renormalize the state manually after each iteration step.Therefore, it is obvious that after certa<strong>in</strong> time τ, the contribution of higher stateswill be negligible, and thus we will arrive at the result given by Eq. (A.2). Asimilar reason<strong>in</strong>g is valid also for nonl<strong>in</strong>ear systems, hence <strong>in</strong> the case of the GPequation we apply the transformation (A.1) and use imag<strong>in</strong>ary-time propagation tof<strong>in</strong>d the stationary states. From the numerical side, imag<strong>in</strong>ary-time and real-timepropagation can be implemented <strong>in</strong> a very similar manner, and we briefly presentonly the details of numerical implementation of the real-time propagation.To simplify the expression for the Laplacian <strong>in</strong> a spherically-symmetric case, wefirst apply a commonly used rescal<strong>in</strong>g:φ(r, t) =ψ(r, t)r, (A.4)and transform the spherically-symmetric GP equation <strong>in</strong>to its simplified form:∂φ(r, t)i∂t[= − 1 ∂ 22 ∂r + 1 2 2 r2 + g∣φ(r, t)r] ∣ φ(r, t) . (A.5)∣2Next, we split the propagation <strong>in</strong>to two steps, which correspond to the two parts ofEq. (A.5):∂φ(r, t)i∂t∂φ(r, t)i∂t=[12 r2 + g∂ 2∣φ(r, t)r= − 1 2 ∂r2φ(r, t) .(A.7)] ∣ φ(r, t) , (A.6)∣2This is the split-step approximation, valid for the short propagation time. It ismotivated by a possibility to treat each of the two previous equations <strong>in</strong> a speciallysuited way: the first equation deals with the part of the Hamiltonian which isdiagonal <strong>in</strong> the coord<strong>in</strong>ate space, while the second equation considers the k<strong>in</strong>eticterm. To perform the time discretization, we <strong>in</strong>troduce the <strong>in</strong>dex n, which countsthe time slices φ(r, t) ≡ φ n (r), where t = nε, and ε is a discrete time-step of thepropagation. Accord<strong>in</strong>g to Eqs. (A.6) and (A.7), there are two different steps to beperformed for one time-step, <strong>in</strong> the propagation from t to t + ε. Correspond<strong>in</strong>gly,we <strong>in</strong>troduce the notation φ n+1/2 (r), which is the value of the wavefunction after the138


first part of the propagation accord<strong>in</strong>g to Eq. (A.6), while after the additional stepaccord<strong>in</strong>g to Eq. (A.7), the propagation to t+ǫ is f<strong>in</strong>ished, and a new wave functionφ n+1 (r) is calculated. F<strong>in</strong>ally, we rewrite Eqs. (A.6) and (A.7) <strong>in</strong> the approximatediscretized form:φ n+1/2 (r) = e −iε „12 r2 +g˛ φn (r) ˛r˛2« φ n (r), (A.8)i 1 (φ n+1 (r) − φ n+1/2 (r) ) = − 1 ∂ 2 φ n+1/2− 1 ∂ 2 φ n+1. (A.9)ε4 ∂r 2 4 ∂r 2In the last equation, we have used the semi-implicit Crank-Nicolson method [137],which is highly accurate, robust and <strong>in</strong>expensive method for solv<strong>in</strong>g general diffusionequations. The accuracy of the approximation is ε 2 .To reduce the obta<strong>in</strong>ed differential equations to the algebraic form, we additionallyperform space discretization with the discretization step h. To this end,we <strong>in</strong>troduce <strong>in</strong>dex i, which takes values from 0 to i max − 1, and approximate thesecond-order spatial derivative <strong>in</strong> the standard way:∂ 2 φ n→ 1 ( )φn∂r 2 4h 2 i+1 − 2φ n i + φ n i−1 , (A.10)with the error of the order of h 2 . As a result, we obta<strong>in</strong> a tridiagonal system ofequations:where−Aφ n+1i+1 + Bφn+1 i − Aφ n+1i−1 = δ i ,(A.11)δ i = (ε/4h 2 )φ n+1/2i+1 + ( 1 − ε/2h 2) φ n+1/2i +(ε/4h 2 )φ n+1/2i−1 , A = ε/4h 2 , B = 1+ε/2h 2 .A solution of a tridiagonal system of equations can be cast <strong>in</strong> the form:φ n+1i+1 = α iφ n+1i + β i , (A.12)and from this Ansatz we f<strong>in</strong>d the recursive relations for the solution:α i−1 =AB − Aα i,β i−1 = δ i + Aβ iB − Aα i.(A.13)From the boundary condition φ n i max= φ n+1/2i max, we derive <strong>in</strong>itial values for α and β:α imax−1 = 0 and β imax−1 = φ n+1/2i max, which we use to solve the recursive equations139


(A.13). With another boundary condition φ n 0 = 0, we f<strong>in</strong>ally solve Eq. (A.12). Dueto the tridiagonal form of the above system of equations, the complexity of thealgorithm is proportional to the number of discretization po<strong>in</strong>ts.Once the ground-state solution φ(r) is calculated us<strong>in</strong>g the imag<strong>in</strong>ary-time versionof the above algorithm, the correspond<strong>in</strong>g values of the chemical potential µand the energy E of the system can be calculated as:∫ [∞( ) rµ = 4π dr φ 2 2(r)0 2 + gφ2 (r) + 1 ( ) ] 2 ∂φ(r),2 ∂r∫ [∞( rE = 4π dr φ 2 2(r)0 2 + g )2 φ2 (r) + 1 ( ) ] 2 ∂φ(r).2 ∂rAn example of the convergence of the value of the chemical potential as a functionof a propagated imag<strong>in</strong>ary-time is shown <strong>in</strong> Fig. A.1. We see the saturation of thechemical-potential value for a long time of propagation.4035g = 627.4g = 3137.13025µ201510500 0.5 1 1.5 2 2.5 3τFigure A.1: The value of the chemical potential <strong>in</strong> units of ω as a function of theimag<strong>in</strong>ary time of propagation τ, <strong>in</strong> units of ω −1 , for two different dimensionless<strong>in</strong>teraction strengths g = 4πNa/l, calculated with the discretization parametersǫ = 10 −4 and h = 10 −2 . We perform the calculation for the spherically symmetrictrap Mω 2 /2, with the characteristic length scale l = √ /Mω. We see that for longtimes of propagation, the value of the chemical potential has converged toward thef<strong>in</strong>al result µ ≈ 7.24836, for g = 627.4, and µ ≈ 13.5534 for g = 3137.1.140


This algorithm can be straightforwardly extended to higher-dimensional cases,as expla<strong>in</strong>ed <strong>in</strong> detail <strong>in</strong> Ref. [87].141


Appendix B Time-dependent variational analysisFor completeness, here we present details of the time-dependent variational analysisused <strong>in</strong> Chapter 5, which was orig<strong>in</strong>ally <strong>in</strong>troduced <strong>in</strong> Refs. [104, 105].We start from the Lagrangian (4.16), assum<strong>in</strong>g a time-dependent <strong>in</strong>teractiong(t). For the Gaussian variational Ansatz (5.1), we calculateψ G∂ψG∗∂t∂ψ G∂σ− ψ G∗∂ψG∂t= −2 i N(t) 2 ∑σ=x,y,z( )∂ψ G∗ (σ − σ2∂σ = 0 )N(t)2 + (ϕu 4 σ + 2σφ σ ) 2 expσ() (σ ˙ϕ σ + σ 2 ˙φσ exp − ∑(− ∑σ=x,y,zσ=x,y,z)(σ − σ 0 ) 2,u 2 σ)(σ − σ 0 ) 2,where we have <strong>in</strong>troduced the notation N(t) 2 = (π 3/2 u x (t)u y (t)u z (t)) −1 , and σ ∈{x, y, z}. Us<strong>in</strong>g the follow<strong>in</strong>g Gaussian <strong>in</strong>tegralsN(t) 2 ∫N(t) 2 ∫N(t) 4 ∫d⃗r σ expd⃗r σ 2 expd⃗r exp(((we calculate the GP LagrangianL GP= ∑σ=x,y,z+ 22M+ ∑σ=x,y,z− ∑σ=x,y,z− ∑σ=x,y,z− ∑σ=x,y,z)(σ − σ 0 ) 2u 2 σ)(σ − σ 0 ) 2u 2 σ)2(σ − σ 0 ) 2u 2 σ(σ 0 ˙ϕ σ + 1 )( )u22 σ + 2σ02 ˙φσ∑σ=x,y,zMω 2 σ2( 12u 2 σ= u σ ,= 1 2=u 2 σ( )u2σ + 2σ02 ,1(2π) 3/2 u x u y u z,+ ϕ 2 σ + 4ϕ σφ σ σ 0 + 2φ 2 σ u2 σ + 4φ2 σ σ2 0( ) u2σ2 + σ2 0+ g(t)21.(2π) 3/2 u x u y u z)142


Euler-Lagrange equations of motion for variational parameters q ∈ {ϕ σ , σ 0 , φ σ , u σ }have the formand for the above Lagrangian read:d ∂Ldt ∂ ˙q − ∂L∂q = 0 , ˙φ σ u σ − 22Mϕ σ − M ˙σ 0 + 2φ σ σ 0 = 0, (B.1) ˙ϕ σ + 2σ 0 ˙φσ + 22M ϕ σφ σ + 42M φ2 σσ 0 + Mω 2 σσ 0 = 0 (B.2)1u 3 σφ σ − M 2˙u σu σ= 0, (B.3)+ 22M φ2 σu σ + Mω2 σ2 u σ − g(t)2(2π) 3/2 1u x u y u z u σ= 0. (B.4)Eqs. (B.1) and (B.3) give the explicit relation between ϕ σ and σ 0 , and between φ σand u σ . By <strong>in</strong>sert<strong>in</strong>g Eqs. (B.1) and (B.3) <strong>in</strong>to Eqs. (B.2) and (B.4), we obta<strong>in</strong>variational equations which we will refer to as a Gaussian approximation:¨σ 0 (t) + λ 2 σ σ 0 = 0 , (B.5)ü σ (t) + λ 2 σ u σ(t) − 1u σ (t) − P(t)3 u σ (t)u x (t)u y (t)u z (t)= 0 . (B.6)We also note that solutions of Eqs. (B.5) and (B.6) for σ 0 (t) and u σ (t) can be afterwards<strong>in</strong>serted <strong>in</strong>to Eqs. (B.1) and (B.3) to obta<strong>in</strong> time-evolution of the phaseparameters ϕ σ and φ σ . The last two equations are given <strong>in</strong> the dimensionless form:we choose a convenient frequency scale ω (for example, the external trap frequency<strong>in</strong> one of the spatial directions) and express all lengths <strong>in</strong> the units of the characteristicharmonic oscillator length l = √ /Mω, time <strong>in</strong> units of ω −1 , and externalfrequencies <strong>in</strong> units of ω: λ σ = ω σ /ω, σ ∈ {x, y, z}. The dimensionless <strong>in</strong>teractionparameter P(t) is given byP(t) =√g(t) 2(2π) 3/2 ωl = 3 π N a(t) .l143


List of papers by Ivana VidanovićThis <strong>thesis</strong> is based on the follow<strong>in</strong>g publications:1. I. Vidanović, A. Balaž, H. Al-Jibbouri, and A. Pelster, Nonl<strong>in</strong>ear Bose-E<strong>in</strong>ste<strong>in</strong>condensateDynamics Induced by a Harmonic Modulation of the s-wave Scatter<strong>in</strong>gLength, Phys. Rev. A 84, 013618 (2011), [Chapter 5]2. A. Balaž, I. Vidanović, A. Bogojević, and A. Pelster, Ultra-fast converg<strong>in</strong>gpath-<strong>in</strong>tegral approach for rotat<strong>in</strong>g ideal Bose-E<strong>in</strong>ste<strong>in</strong> condensates, Phys.Lett. A 374, 1539 (2010), [Chapter 3]3. I. Vidanović, A. Bogojević, A. Balaž, and A. Belić, Properties of QuantumSystems Via Diagonalization of Transition Amplitudes. II. Systematic Improvementsof Short-time Propagation, Phys. Rev. E 80, 066706 (2009),[Chapter 2]4. I. Vidanović, A. Bogojević, and A. Belić, Properties of Quantum SystemsVia Diagonalization of Transition Amplitudes. I. Discretization Effects, Phys.Rev. E 80, 066705 (2009), [Chapter 2]Publications not <strong>in</strong>cluded <strong>in</strong> the <strong>thesis</strong>:1. A. Balaž, I. Vidanović, D. Stojiljković, D. Vudragović, A. Belić, and A. Bogojević,SPEEDUP Code for Calculation of Transition Amplitudes Via theEffective Action Approach, Commun. Comput. Phys. 11, 739 (2012).2. I. Vidanović, S. Arsenijević, and S. Elezović-Hadzić, Force-<strong>in</strong>duced Desorptionof Self-avoid<strong>in</strong>g Walks on Sierp<strong>in</strong>ski Gasket Fractals, Eur. Phys. J. B. 81,291 (2011).3. A. Balaž, I. Vidanović, A. Bogojević, A. Belić, and A. Pelster, Fast Converg<strong>in</strong>gPath Integrals for Time-Dependent Potentials: II. Generalization to Many-Body Systems and Real-Time Formalism, J. Stat. Mech. P03005 (2011).144


4. A. Balaž, I. Vidanović, A. Bogojević, A. Belić, and A. Pelster, Fast Converg<strong>in</strong>gPath Integrals for Time-Dependent Potentials: Recursive Calculation of Short-Time Expansion of the Propagator, J. Stat. Mech. P03004 (2011).5. M. V. Milovanović, Th. Jolicoeur, and I. Vidanović, Modified Coulomb GasConstruction of Quantum Hall States from Nonunitary Conformal Field Theories,Phys. Rev. B 80, 155324 (2009).6. A. Balaž, A. Bogojević, I. Vidanović, and A. Pelster, Recursive Schröd<strong>in</strong>gerEquation Approach to Faster Converg<strong>in</strong>g Path Integrals, Phys. Rev. E 79,036701 (2009).7. A. Bogojević, I. Vidanović, A. Balaž, and A. Belić, Fast Convergence of PathIntegrals for Many-Body Systems, Phys. Lett. A 372, 3341 (2008).145


References[1] S. N. Bose, Plancks gesetz und lichtquantenhypothese, Z. Phys. 26, 178(1924). v, x, 1[2] A. E<strong>in</strong>ste<strong>in</strong>, Quantentheorie des e<strong>in</strong>atomigen idealen gases, Sitzungsber.Preuss. Akad. Wiss. p. 261 (1924). v, x, 1[3] M. H. Anderson, J. R. Ensher, M. R. Matthews, C. E. Wieman, and E. A.Cornell, Observation of Bose-E<strong>in</strong>ste<strong>in</strong> condensation <strong>in</strong> a dilute atomic vapor,Science 269, 198 (1995). vi, xi, 2, 7, 11[4] K. B. Davis, M. O. Mewes, M. R. Andrews, N. J. van Druten, D. S. Durfee,D. M. Kurn, and W. Ketterle, Bose-E<strong>in</strong>ste<strong>in</strong> condensation <strong>in</strong> a gas of sodiumatoms, Phys. Rev. Lett. 75, 3969 (1995). vi, xi, 2, 11[5] F. Dalfovo, S. Giorg<strong>in</strong>i, L. P. Pitaevskii, and S. Str<strong>in</strong>gari, Theory of Bose-E<strong>in</strong>ste<strong>in</strong> condensation <strong>in</strong> trapped gases, Rev. Mod. Phys. 71, 463 (1999). vi,xi, 5, 7, 12, 84, 105[6] C. Ch<strong>in</strong>, R. Grimm, P. Julienne, and E. Ties<strong>in</strong>ga, Feshbach resonances <strong>in</strong>ultracold gases, Rev. Mod. Phys. 82, 1225 (2010). vi, xi, 13[7] I. Bloch, J. Dalibard, and W. Zwerger, Many-body physics with ultracoldgases, Rev. Mod. Phys. 80, 885 (2008). vi, vii, xi, xii, 15, 57, 58[8] R. Feynman, Simulat<strong>in</strong>g physics with computers, Int. J. Theor. Phys. 21, 467(1982). vi, xi, 3[9] A. Sethia, S. Sanyal, and Y. S<strong>in</strong>gh, Discretized path <strong>in</strong>tegral method andproperties of a quantum system, J. Chem. Phys. 93, 7268 (1990). vii, xii, 16,20, 21, 23, 33, 55146


REFERENCES[10] A. Bogojević, A. Balaž, and A. Belić, Systematically accelerated convergenceof path <strong>in</strong>tegrals, Phys. Rev. Lett. 94, 180403 (2005). vii, xii, 16, 21, 34, 65[11] A. Bogojević, I. Vidanović, A. Balaž, and A. Belić, Fast convergence of path<strong>in</strong>tegrals for many-body systems, Phys. Lett. A 372, 3341 (2008). vii, xii, 16,21, 34[12] V. Bret<strong>in</strong>, S. Stock, Y. Seur<strong>in</strong>, and J. Dalibard, Fast rotation of a Bose-E<strong>in</strong>ste<strong>in</strong> condensate, Phys. Rev. Lett. 92, 050403 (2004). vii, xii, 17, 47, 49,58, 59, 62, 68, 70, 71[13] Gross E. P., Structure of a quantized vortex <strong>in</strong> boson systems, Nuovo Cimento20, 454 (1961). viii, xiii, 17, 88[14] Pitaevskii L. P., Vortex l<strong>in</strong>es <strong>in</strong> an imperfect Bose gas, Sov. Phys. JETP 13,451 (1961). viii, xiii, 17, 88[15] S. E. Pollack, D. Dries, R. G. Hulet, K. M. F. Magalhães, E. A. L. Henn,E. R. F. Ramos, M. A. Caracanhas, and V. S. Bagnato, Collective excitationof a Bose-E<strong>in</strong>ste<strong>in</strong> condensate by modulation of the atomic scatter<strong>in</strong>g length,Phys. Rev. A 81, 053627 (2010). viii, xiii, 17, 107, 112, 113, 114, 126, 131[16] M. R. Andrews, M.-O. Mewes, N. J. van Druten, D. S. Durfee, D. M. Kurn,and W. Ketterle, Direct, nondestructive observation of a Bose condensate,Science 273, 84 (1996). 2, 11[17] J. Klaers, J. Schmitt, F. Vew<strong>in</strong>ger, and M. Weitz, Bose-E<strong>in</strong>ste<strong>in</strong> condensationof photons <strong>in</strong> an optical microcavity, Nature (London) 468, 545 (2010). 3[18] V. Bagnato, D. E. Pritchard, and D. Kleppner, Bose-E<strong>in</strong>ste<strong>in</strong> condensation<strong>in</strong> an external potential, Phys. Rev. A 35, 4354 (1987). 3[19] C. Pethick and H. Smith, Bose-E<strong>in</strong>ste<strong>in</strong> Condensation <strong>in</strong> Dilute Gases, CambridgeUniversity Press, Cambridge (2008). 3, 5, 6, 8, 13, 17, 57, 59, 86, 88,90, 106[20] W. Ketterle and N. J. van Druten, Bose-E<strong>in</strong>ste<strong>in</strong> condensation of a f<strong>in</strong>itenumber of particles trapped <strong>in</strong> one or three dimensions, Phys. Rev. A 54, 656(1996). 5147


REFERENCES[21] W. Ketterle, D. S. Durfee, and D. M. Stamper-Kurn, Mak<strong>in</strong>g, prob<strong>in</strong>g andunderstand<strong>in</strong>g Bose-E<strong>in</strong>ste<strong>in</strong> condensates, arXiv:cond-mat/9904034 (1999). 8,9, 10, 109[22] H. Metcalf and P. van der Straten, Cool<strong>in</strong>g and trapp<strong>in</strong>g of neutral atoms,Phys. Rep. 244, 203 (1994). 10[23] K. B. Davis, M.-O. Mewes, M. A. Joffe, M. R. Andrews, and W. Ketterle,Evaporative cool<strong>in</strong>g of sodium atoms, Phys. Rev. Lett. 74, 5202 (1995). 10[24] R. Meppel<strong>in</strong>k, R. A. Rozendaal, S. B. Koller, J. M. Vogels, and P. van derStraten, Thermodynamics of Bose-E<strong>in</strong>ste<strong>in</strong>-condensed clouds us<strong>in</strong>g phasecontrastimag<strong>in</strong>g, Phys. Rev. A 81, 053632 (2010). 11, 101[25] T. Lahaye, T. Koch, B. Frohlich, M. Fattori, J. Metz, A. Griesmaier, S. Giovanazzi,and T. Pfau, Strong dipolar effects <strong>in</strong> a quantum ferrofluid, Nature(London) 448, 672 (2007). 12[26] H. Feshbach, Unified theory of nuclear reactions, Ann. Phys. (NY) 5, 357(1958). 13[27] U. Fano, Effects of configuration <strong>in</strong>teraction on <strong>in</strong>tensities and phase shifts,Phys. Rev. 124, 1866 (1961). 13[28] E. Ties<strong>in</strong>ga, B. J. Verhaar, and H. T. C. Stoof, Threshold and resonancephenomena <strong>in</strong> ultracold ground-state collisions, Phys. Rev. A 47, 4114 (1993).13[29] J.T.M. Walraven, Elements of Quantum Gases: Thermodynamic and CollisionalProperties of Trapped Atomic Gases, Les Houches - predoc lectures2010, (unpublished). 13[30] N. N. Bogoliubov, On the theory of superfluidity, J. Phys. (USSR) 11, 23(1947). 14, 15, 104[31] M. Stone, The Physics of Quantum Fields, Spr<strong>in</strong>ger-Verlag, New York (2000).15[32] A. Altland and B. Simons, Condensed Matter Field Theory, Cambridge UniversityPress, Cambridge (2006). 15, 85148


REFERENCES[33] D. A. Huse and E. D. Siggia, The density distribution of a weakly <strong>in</strong>teract<strong>in</strong>gbose gas <strong>in</strong> an external potential, J. Low Temp. Phys. 46, 137 (1982). 15, 95[34] L. Pitaevskii and S. Str<strong>in</strong>gari, Bose-E<strong>in</strong>ste<strong>in</strong> Condensation, Clarendon Press,Oxford (2004). 17, 88[35] A. Sethia, S. Sanyal, and F. Hirata, Eigenstates from the discretized path<strong>in</strong>tegral, Chem. Phys. Lett. 315, 299 (1999). 20[36] A. Sethia, S. Sanyal, and F. Hirata, Quantum dynamics: Path <strong>in</strong>tegral approachto time correlation functions <strong>in</strong> f<strong>in</strong>ite temperature, J. Chem. Phys.114, 5097 (2001). 20[37] S. Sanyal and A. Sethia, Quantum properties of many boson system: Path<strong>in</strong>tegral approach, Chem. Phys. Lett. 404, 192 (2005). 20[38] H. Ishikawa, An accurate method for numerical calculations <strong>in</strong> quantum mechanics,J. Phys. A 35, 4453 (2002). 21[39] H. Shao and Z. Wang, Numerical solutions of the time-dependent Schröd<strong>in</strong>gerequation: Reduction of the error due to space discretization, Phys. Rev. E 79,056705 (2009). 21[40] S. A. Ch<strong>in</strong>, S. Janecek, and E. Krotscheck, An arbitrary order diffusion algorithmfor solv<strong>in</strong>g Schröd<strong>in</strong>ger equations, Comput. Phys. Commun. 180, 1700(2009). 21[41] S. Janecek and E. Krotscheck, A fast and simple program for solv<strong>in</strong>g localSchröd<strong>in</strong>ger equations <strong>in</strong> two and three dimensions, Comput. Phys. Commun.178, 835 (2008). 21[42] M. Aich<strong>in</strong>ger, S.A. Ch<strong>in</strong>, and E. Krotscheck, Fourth-order algorithms forsolv<strong>in</strong>g local Schröd<strong>in</strong>ger equations <strong>in</strong> a strong magnetic field, Comput. Phys.Commun. 171, 197 (2005). 21[43] A. Balaž, O ubrzavanju konvergencije funkcionalnih <strong>in</strong>tegrala, <strong>PhD</strong> <strong>thesis</strong>,Faculty of Physics, University of Belgrade (2008). 21[44] A. Balaž, A. Bogojević, I. Vidanović, and A. Pelster, Recursive Schröd<strong>in</strong>gerequation approach to faster converg<strong>in</strong>g path <strong>in</strong>tegrals, Phys. Rev. E 79, 036701(2009). 21, 26, 34, 36, 37, 55, 65149


REFERENCES[45] A. Bogojević, A. Balaž, and A. Belić, Systematic speedup of path <strong>in</strong>tegrals ofa generic N -fold discretized theory, Phys. Rev. B 72, 064302 (2005). 21, 34[46] A. Bogojević, A. Balaž, and A. Belić, Generalization of Euler’s summationformula to path <strong>in</strong>tegrals, Phys. Lett. A 344, 84 (2005). 21, 34[47] T. L. Beck, Real-space mesh techniques <strong>in</strong> density-functional theory, Rev.Mod. Phys. 72, 1041 (2000). 22[48] M. A. Mart<strong>in</strong>-Delgado, G. Sierra, and R. M. Noack, The density matrixrenormalization group applied to s<strong>in</strong>gle-particle quantum mechanics, J. Phys.A 32, 6079 (1999). 23, 56[49] J. R. Chelikowsky, N. Troullier, K. Wu, and Y. Saad, Higher-order f<strong>in</strong>itedifferencepseudopotential method: An application to diatomic molecules,Phys. Rev. B 50, 11355 (1994). 23[50] P. Maragakis, J. Soler, and E. Kaxiras, Variational f<strong>in</strong>ite-difference representationof the k<strong>in</strong>etic energy operator, Phys. Rev. B 64, 193101 (2001).23[51] C.-K. Skylaris, A. A. Mostofi, P. D. Haynes, O. Diéguez, and M. C.Payne, Nonorthogonal generalized Wannier function pseudopotential planewavemethod, Phys. Rev. B 66, 035119 (2002). 23[52] H. Kle<strong>in</strong>ert, Path Integrals <strong>in</strong> Quantum Mechanics, Statistics, PolymerPhysics, and F<strong>in</strong>ancial Markets, World Scientific, S<strong>in</strong>gapore (2006). 27, 43[53] W. Janke and H. Kle<strong>in</strong>ert, Summ<strong>in</strong>g paths for a particle <strong>in</strong> a box, Lett. NuovoCimento 25, 297 (1979). 27[54] Mathematica symbolic calculation software packagehttp://www.wolfram.com/mathematica. 30, 40, 120, 122, 129[55] G. Barton, A. J. Bray, and A. J. McKane, The <strong>in</strong>fluence of distant boundarieson quantum mechanical energy levels, Am. J. Phys. 58, 751 (1990). 31, 38[56] D. H. Berman, Boundary effects <strong>in</strong> quantum mechanics, Am. J. Phys. 59,937 (1991). 31, 38150


REFERENCES[57] D. M. Ceperley, Path <strong>in</strong>tegrals <strong>in</strong> the theory of condensed helium, Rev. Mod.Phys. 67, 279 (1995). 34[58] M. Bon<strong>in</strong>segni, N. Prokof’ev, and B. Svistunov, Worm algorithm forcont<strong>in</strong>uous-space path <strong>in</strong>tegral Monte Carlo simulations, Phys. Rev. Lett.96, 070601 (2006). 34[59] A. Bogojević, A. Balaž, and A. Belić, Asymptotic properties of path <strong>in</strong>tegralideals, Phys. Rev. E 72, 036128 (2005). 36[60] SPEEDUP C and SCL BEC Mathematica codes,http://www.scl.rs/speedup/. 37, 41, 44, 49, 64, 123, 129[61] W. Janke and H. Kle<strong>in</strong>ert, Convergent strong-coupl<strong>in</strong>g expansions from divergentweak-coupl<strong>in</strong>g perturbation theory, Phys. Rev. Lett. 75, 2787 (1995).37, 39[62] LAPACK – L<strong>in</strong>ear Algebra PACKage, http://www.netlib.org/lapack/. 41,49[63] I. S. Gradshteyn and I. M. Ryzhik, Table of <strong>in</strong>tegrals, series, and products,Academic Press, San Diego (2000). 43, 95[64] S. Stock, B. Battelier, V. Bret<strong>in</strong>, Z. Hadžibabić, and J. Dalibard, Bose-E<strong>in</strong>ste<strong>in</strong>condensates <strong>in</strong> fast rotation, Las. Phys. Lett. 2, 275 (2005). 47, 49[65] S. Kl<strong>in</strong>g and A. Pelster, Thermodynamical properties of a rotat<strong>in</strong>g ideal Bosegas, Phys. Rev. A 76, 023609 (2007). 47, 59, 68, 71, 73, 74, 75, 76, 80[66] T. H. Seligman, J. J. M. Verbaarschot, and M. R. Zirnbauer, Quantum spectraand transition from regular to chaotic classical motion, Phys. Rev. Lett. 53,215 (1984). 53[67] U. Schollwöck, The density-matrix renormalization group, Rev. Mod. Phys.77, 259 (2005). 56[68] A. L. Fetter, Rotat<strong>in</strong>g trapped Bose-E<strong>in</strong>ste<strong>in</strong> condensates, Rev. Mod. Phys.81, 647 (2009). 57[69] K. W. Madison, F. Chevy, W. Wohlleben, and J. Dalibard, Vortex formation<strong>in</strong> a stirred Bose-E<strong>in</strong>ste<strong>in</strong> condensate, Phys. Rev. Lett. 84, 806 (2000). 57151


REFERENCES[70] M. R. Matthews, B. P. Anderson, P. C. Haljan, D. S. Hall, C. E. Wieman,and E. A. Cornell, Vortices <strong>in</strong> a Bose-E<strong>in</strong>ste<strong>in</strong> condensate, Phys. Rev. Lett.83, 2498 (1999). 57[71] A. A. Abrikosov, On the magnetic properties of superconductors of the secondgroup, Sov. Phys. JETP 5, 1174 (1957). 57[72] J. R. Abo-Shaeer, C. Raman, J. M. Vogels, and W. Ketterle, Observation ofvortex lattices <strong>in</strong> Bose-E<strong>in</strong>ste<strong>in</strong> condensates, Science 292, 476 (2001). 57[73] S. Viefers, Quantum Hall physics <strong>in</strong> rotat<strong>in</strong>g BoseE<strong>in</strong>ste<strong>in</strong> condensates, J.Phys.- Condens. Matter 20, 123202 (2008). 57[74] A. L. Fetter, Rotat<strong>in</strong>g vortex lattice <strong>in</strong> a Bose-E<strong>in</strong>ste<strong>in</strong> condensate trapped<strong>in</strong> comb<strong>in</strong>ed quadratic and quartic radial potentials, Phys. Rev. A 64, 063608(2001). 58[75] K. Howe, A. R. P. Lima, and A. Pelster, Rotat<strong>in</strong>g Fermi gases <strong>in</strong> an anharmonictrap, Eur. Phys. J. D 54, 667 (2009). 83[76] M. Holzmann, W. Krauth, and M. Naraschewski, Precision Monte Carlo testof the Hartree-Fock approximation for a trapped Bose gas, Phys. Rev. A 59,2956 (1999). 84, 97[77] W. Krauth, Quantum Monte Carlo calculations for a large number of bosons<strong>in</strong> a harmonic trap, Phys. Rev. Lett. 77, 3695 (1996). 84[78] S. Giorg<strong>in</strong>i, L. Pitaevskii, and S. Str<strong>in</strong>gari, Thermodynamics of a trappedBose-condensed gas, J. Low Temp. Phys. 109, 309 (1997). 84, 87, 101[79] V. I. Yukalov, Pr<strong>in</strong>cipal problems <strong>in</strong> Bose-E<strong>in</strong>ste<strong>in</strong> condensation of dilutegases, Las. Phys. Lett. 1, 435 (2004). 85[80] N. Tammuz, R. P. Smith, R. L. D. Campbell, S. Beattie, S. Moulder, J. Dalibard,and Z. Hadžibabić, Can a Bose gas be saturated?, Phys. Rev. Lett.106, 230401 (2011). 85, 102[81] R. P. Smith, R. L. D. Campbell, N. Tammuz, and Z. Hadžibabić, Effects of<strong>in</strong>teractions on the critical temperature of a trapped Bose gas, Phys. Rev.Lett. 106, 250403 (2011). 85, 102152


REFERENCES[82] N. P. Proukakis and B. Jackson, F<strong>in</strong>ite-temperature models of BoseE<strong>in</strong>ste<strong>in</strong>condensation, J. Phys. B 41, 203002 (2008). 86[83] A. Griff<strong>in</strong>, Conserv<strong>in</strong>g and gapless approximations for an <strong>in</strong>homogeneous Bosegas at f<strong>in</strong>ite temperatures, Phys. Rev. B 53, 9341 (1996). 86[84] V. I. Yukalov and H. Kle<strong>in</strong>ert, Gapless Hartree-Fock-Bogoliubov approximationfor Bose gases, Phys. Rev. A 73, 063612 (2006). 87[85] Y. S. Kivshar and B. A. Malomed, Dynamics of solitons <strong>in</strong> nearly <strong>in</strong>tegrablesystems, Rev. Mod. Phys. 61, 763 (1989). 88, 106[86] M. Belić, N. Petrović, W.-P. Zhong, R.-H. Xie, and G. Chen, Analytical lightbullet solutions to the generalized (3 + 1)-dimensional nonl<strong>in</strong>ear Schröd<strong>in</strong>gerequation, Phys. Rev. Lett. 101, 123904 (2008). 88, 106[87] P. Muruganandam and S. K. Adhikari, Fortran programs for the timedependentGrossPitaevskii equation <strong>in</strong> a fully anisotropic trap, Comput. Phys.Commun. 180, 1888 (2009). 89, 116, 125, 137, 141[88] M. Naraschewski and D. M. Stamper-Kurn, Analytical description of atrapped semi-ideal Bose gas at f<strong>in</strong>ite temperature, Phys. Rev. A 58, 2423(1998). 91, 92, 93[89] F. Gerbier, J. H. Thywissen, S. Richard, M. Hugbart, P. Bouyer, and A. Aspect,Experimental study of the thermodynamics of an <strong>in</strong>teract<strong>in</strong>g trappedBose-E<strong>in</strong>ste<strong>in</strong> condensed gas, Phys. Rev. A 70, 013607 (2004). 91, 92, 93, 101[90] S. Giorg<strong>in</strong>i, L. P. Pitaevskii, and S. Str<strong>in</strong>gari, Condensate fraction and criticaltemperature of a trapped <strong>in</strong>teract<strong>in</strong>g Bose gas, Phys. Rev. A 54, R4633 (1996).99[91] S. Grossmann and M. Holthaus, On Bose-E<strong>in</strong>ste<strong>in</strong> condensation <strong>in</strong> harmonictraps, Phys. Lett. A 208, 188 (1995). 100[92] T. Haugset, H. Haugerud, and J. O. Andersen, Bose-E<strong>in</strong>ste<strong>in</strong> condensation <strong>in</strong>anisotropic harmonic traps, Phys. Rev. A 55, 2922 (1997). 100[93] J. R. Ensher, D. S. J<strong>in</strong>, M. R. Matthews, C. E. Wieman, and E. A. Cornell,Bose-E<strong>in</strong>ste<strong>in</strong> condensation <strong>in</strong> a dilute gas: Measurement of energy andground-state occupation, Phys. Rev. Lett. 77, 4984 (1996). 100153


REFERENCES[94] M.-O. Mewes, M. R. Andrews, N. J. van Druten, D. M. Kurn, D. S. Durfee, andW. Ketterle, Bose-E<strong>in</strong>ste<strong>in</strong> condensation <strong>in</strong> a tightly conf<strong>in</strong><strong>in</strong>g dc magnetictrap, Phys. Rev. Lett. 77, 416 (1996). 100[95] A. Griff<strong>in</strong>, T. Nikuni, and E. Zaremba, Bose-Condensed Gases at F<strong>in</strong>ite Temperatures,Cambridge University Press, Cambridge (2009). 102[96] L. D. Landau and E. M. Lifshitz, Course of Theoretical Physics, Statisticalphysics, Pergamon press, Oxford (1980). 104[97] D. S. J<strong>in</strong>, J. R. Ensher, M. R. Matthews, C. E. Wieman, and E. A. Cornell,Collective excitations of a Bose-E<strong>in</strong>ste<strong>in</strong> condensate <strong>in</strong> a dilute gas, Phys. Rev.Lett. 77, 420 (1996). 105[98] M.-O. Mewes, M. R. Andrews, N. J. van Druten, D. M. Kurn, D. S. Durfee,C. G. Townsend, and W. Ketterle, Collective excitations of a Bose-E<strong>in</strong>ste<strong>in</strong>condensate <strong>in</strong> a magnetic trap, Phys. Rev. Lett. 77, 988 (1996). 105[99] F. Chevy, V. Bret<strong>in</strong>, P. Rosenbusch, K. W. Madison, and J. Dalibard, Transversebreath<strong>in</strong>g mode of an elongated Bose-E<strong>in</strong>ste<strong>in</strong> condensate, Phys. Rev.Lett. 88, 250402 (2002). 105, 111[100] D. M. Stamper-Kurn, H.-J. Miesner, S. Inouye, M. R. Andrews, and W. Ketterle,Collisionless and hydrodynamic excitations of a Bose-E<strong>in</strong>ste<strong>in</strong> condensate,Phys. Rev. Lett. 81, 500 (1998). 105, 111[101] H. Ott, J. Fortágh, S. Kraft, A. Günther, D. Komma, and C. Zimmermann,Nonl<strong>in</strong>ear dynamics of a Bose-E<strong>in</strong>ste<strong>in</strong> condensate <strong>in</strong> a magnetic waveguide,Phys. Rev. Lett. 91, 040402 (2003). 105, 111, 132[102] M. Edwards, P. A. Ruprecht, K. Burnett, R. J. Dodd, and C. W. Clark,Collective excitations of atomic Bose-E<strong>in</strong>ste<strong>in</strong> condensates, Phys. Rev. Lett.77, 1671 (1996). 106, 111[103] S. Str<strong>in</strong>gari, Collective excitations of a trapped Bose-condensed gas, Phys.Rev. Lett. 77, 2360 (1996). 106, 111[104] V. M. Pérez-García, H. Mich<strong>in</strong>el, J. I. Cirac, M. Lewenste<strong>in</strong>, and P. Zoller,Low energy excitations of a Bose-E<strong>in</strong>ste<strong>in</strong> condensate: A time-dependent variationalanalysis, Phys. Rev. Lett. 77, 5320 (1996). 106, 107, 108, 111, 142154


REFERENCES[105] V. M. Pérez-García, H. Mich<strong>in</strong>el, J. I. Cirac, M. Lewenste<strong>in</strong>, and P. Zoller,Dynamics of Bose-E<strong>in</strong>ste<strong>in</strong> condensates: Variational solutions of the Gross-Pitaevskii equations, Phys. Rev. A 56, 1424 (1997). 106, 107, 108, 111, 142[106] C. Becker, S. Stellmer, P. Soltan-Panahi, S. Dorscher, M. Baumert, E.-M.Richter, J. Kronjager, K. Bongs, and K. Sengstock, Oscillations and <strong>in</strong>teractionsof dark and dark-bright solitons <strong>in</strong> Bose-E<strong>in</strong>ste<strong>in</strong> condensates, Nat.Phys. 4, 496 (2008). 106[107] C. Hamner, J. J. Chang, P. Engels, and M. A. Hoefer, Generation of darkbrightsoliton tra<strong>in</strong>s <strong>in</strong> superfluid-superfluid counterflow, Phys. Rev. Lett. 106,065302 (2011). 106[108] M. Faraday, On a peculiar class of acoustical figures; and on certa<strong>in</strong> formsassumed by a group of particles upon vibrat<strong>in</strong>g elastic surfaces, Philos. Trans.R. Soc. London 121, 299 (1831). 106[109] M. C. Cross and P. C. Hohenberg, Pattern formation outside of equilibrium,Rev. Mod. Phys. 65, 851 (1993). 106[110] J. P. Gollub and J. S. Langer, Pattern formation <strong>in</strong> nonequilibrium physics,Rev. Mod. Phys. 71, S396 (1999). 106[111] P. Engels, C. Atherton, and M. A. Hoefer, Observation of Faraday waves <strong>in</strong> aBose-E<strong>in</strong>ste<strong>in</strong> wondensate, Phys. Rev. Lett. 98, 095301 (2007). 106, 115[112] E. A. L. Henn, J. A. Seman, G. Roati, K. M. F. Magalhães, and V. S. Bagnato,Emergence of turbulence <strong>in</strong> an oscillat<strong>in</strong>g Bose-E<strong>in</strong>ste<strong>in</strong> condensate, Phys.Rev. Lett. 103, 045301 (2009). 106[113] S. K. Adhikari, Resonance <strong>in</strong> BoseE<strong>in</strong>ste<strong>in</strong> condensate oscillation from a periodicvariation <strong>in</strong> scatter<strong>in</strong>g length, J. Phys. B 36, 1109 (2003). 107, 114[114] F. K. Abdullaev, R. M. Galimzyanov, M. Brtka, and R. A. Kraenkel, Resonances<strong>in</strong> a trapped 3D BoseE<strong>in</strong>ste<strong>in</strong> condensate under periodically vary<strong>in</strong>gatomic scatter<strong>in</strong>g length, J. Phys. B 37, 3535 (2004). 107, 114[115] F. K. Abdullaev and J. Garnier, Collective oscillations of one-dimensionalBose-E<strong>in</strong>ste<strong>in</strong> gas <strong>in</strong> a time-vary<strong>in</strong>g trap potential and atomic scatter<strong>in</strong>glength, Phys. Rev. A 70, 053604 (2004). 107, 114155


REFERENCES[116] K. Staliunas, S. Longhi, and G. J. de Valcárcel, Faraday patterns <strong>in</strong> Bose-E<strong>in</strong>ste<strong>in</strong> condensates, Phys. Rev. Lett. 89, 210406 (2002). 107, 114, 115[117] F. Dalfovo, C. M<strong>in</strong>niti, and L. P. Pitaevskii, Frequency shift and mode coupl<strong>in</strong>g<strong>in</strong> the nonl<strong>in</strong>ear dynamics of a Bose-condensed gas, Phys. Rev. A 56,4855 (1997). 111, 114[118] J. J. García-Ripoll, V. M. Pérez-García, and P. Torres, Extended parametricresonances <strong>in</strong> nonl<strong>in</strong>ear Schröd<strong>in</strong>ger systems, Phys. Rev. Lett. 83, 1715 (1999).111, 125, 126[119] J. J. G. Ripoll and V. M. Pérez-García, Barrier resonances <strong>in</strong> Bose-E<strong>in</strong>ste<strong>in</strong>condensation, Phys. Rev. A 59, 2220 (1999). 111[120] J.-K. Xue, G.-Q. Li, A.-X. Zhang, and P. Peng, Nonl<strong>in</strong>ear mode coupl<strong>in</strong>g andresonant excitations <strong>in</strong> two-component Bose-E<strong>in</strong>ste<strong>in</strong> condensates, Phys. Rev.E 77, 016606 (2008). 111[121] H. Pu and N. P. Bigelow, Collective excitations, metastability, and nonl<strong>in</strong>earresponse of a trapped two-species Bose-E<strong>in</strong>ste<strong>in</strong> condensate, Phys. Rev. Lett.80, 1134 (1998). 111[122] S. E. Pollack, D. Dries, M. Junker, Y. P. Chen, T. A. Corcovilos, and R. G.Hulet, Extreme tunability of <strong>in</strong>teractions <strong>in</strong> a 7 Li Bose-E<strong>in</strong>ste<strong>in</strong> condensate,Phys. Rev. Lett. 102, 090402 (2009). 111, 112[123] V. I. Yukalov, E. P. Yukalova, and V. S. Bagnato, Bose systems <strong>in</strong> spatiallyrandom or time-vary<strong>in</strong>g potentials, Las. Phys. 19, 686 (2009). 114[124] N. N. Bogoliubov and Y. A. Mitropolsky, Asymptotic Methods <strong>in</strong> the Theoryof Non-L<strong>in</strong>ear Oscillations, Gordon&Breach, New York (1961). 114, 121, 125[125] Y. Zhou, W. Wen, and G. Huang, Frequency shift and mode coupl<strong>in</strong>g of thecollective modes of superfluid Fermi gases <strong>in</strong> the BCS-BEC crossover, Phys.Rev. B 77, 104527 (2008). 114[126] A. Pelster, H. Kle<strong>in</strong>ert, and M. Schanz, High-order variational calculation forthe frequency of time-periodic solutions, Phys. Rev. E 67, 016604 (2003). 114,121156


REFERENCES[127] R. Mickens, Introduction to Nonl<strong>in</strong>ear Oscillations, Cambridge UniversityPress, Cambridge (1981). 114, 121[128] N. M<strong>in</strong>orsky, Nonl<strong>in</strong>ear Oscillation, Van Nostrand, Pr<strong>in</strong>ceton (1962). 114,121[129] A. I. Nicol<strong>in</strong>, R. Carretero-González, and P. G. Kevrekidis, Faraday waves <strong>in</strong>Bose-E<strong>in</strong>ste<strong>in</strong> condensates, Phys. Rev. A 76, 063609 (2007). 115[130] D. Leibfried, R. Blatt, C. Monroe, and D. W<strong>in</strong>eland, Quantum dynamics ofs<strong>in</strong>gle trapped ions, Rev. Mod. Phys. 75, 281 (2003). 120[131] L. P. Pitaevskii, Dynamics of collapse of a conf<strong>in</strong>ed Bose gas, Phys. Lett. A221, 14 (1996). 126[132] L. P. Pitaevskii and A. Rosch, Breath<strong>in</strong>g modes and hidden symmetry oftrapped atoms <strong>in</strong> two dimensions, Phys. Rev. A 55, R853 (1997). 126[133] L. Salasnich, A. Parola, and L. Reatto, Dimensional reduction <strong>in</strong> Bose-E<strong>in</strong>ste<strong>in</strong>-condensed alkali-metal vapors, Phys. Rev. A 69, 045601 (2004). 131[134] L. Pitaevskii and S. Str<strong>in</strong>gari, Elementary excitations <strong>in</strong> trapped Bose-E<strong>in</strong>ste<strong>in</strong>condensed gases beyond the mean-field approximation, Phys. Rev. Lett. 81,4541 (1998). 133[135] E. Braaten and J. Pearson, Semiclassical corrections to the oscillation frequenciesof a trapped Bose-E<strong>in</strong>ste<strong>in</strong> condensate, Phys. Rev. Lett. 82, 255 (1999).133[136] A. R. P. Lima and A. Pelster, Quantum fluctuations <strong>in</strong> dipolar bose gases,Phys. Rev. A 84, 041604 (2011). 133[137] W. H. Press, S. A. Teukolsky, W. T. Vetterl<strong>in</strong>g, and B. P. Flannery, NumericalRecipes <strong>in</strong> C, Cambridge University Press, Cambridge (1997). 139157


CURRICULUM VITAE - Ivana VidanovićEducation• (2007-2011) <strong>PhD</strong> <strong>in</strong> Condensed Matter Physics, Faculty of Physics,University of Belgrade• (2001-2006) BSc <strong>in</strong> Physics, Faculty of Physics, University of BelgradeMajor: Theoretical Physics• (1997-2001) Mathematical High School, Belgrade, SerbiaEmployment and research projects• (2011-2012) Serbian-German bilateral research project “Numerical andAnalytical Investigation of Ultracold Bose Gases <strong>in</strong> Disordered Potentials”,funded by the Serbian M<strong>in</strong>istry of Education and Science and DAAD• (2011-2014) National research project ON171017 “Model<strong>in</strong>g and NumericalSimulations of Complex Physical Systems”, funded by the Serbian M<strong>in</strong>istry ofEducation and Science• (2009-2010) Serbian-German bilateral research project “Fast Converg<strong>in</strong>g PathIntegral Approach to Bose-E<strong>in</strong>ste<strong>in</strong> Condensation”, funded by the SerbianM<strong>in</strong>istry of Science and DAAD• (2008-present) Employed at the Scientific Comput<strong>in</strong>g Laboratory of the Instituteof Physics Belgrade, Serbia• (2007-2010) National research project OI141035 “Model<strong>in</strong>g and numericalsimulations of complex physical systems” funded by the Serbian M<strong>in</strong>istry ofScience• (2006-2009) FP6 project of the European Commission “CX-CMCS: EU Centreof Excellence for Computer Model<strong>in</strong>g of Complex Systems”158


Collaboration with other <strong>in</strong>stitutions• (May 2011) Visit to Condensed Matter Group, MPI-PKS, Dresden, Germany,host Dr. Masud Haque• (2009-2010) Several visits to Dr. Axel Pelster at the Free University of Berl<strong>in</strong>,University of Duisburg-Essen, and Potsdam University• (June-September 2005) Summer student at the Paul Drude Institute, Berl<strong>in</strong>,GermanyAttended workshops and conferences• March 2011 DPG-2011 Conference, Dresden, Germany• August 2010 Quo Vadis BEC? conference, MPI-PKS Dresden, Germany• July 2010 Many-Body Physics with Ultracold Gases, Les-Houches, France• September 2009 Arnold Sommerfeld Summer School on Condensed MatterPhysics with Ultracold Quantum Gases, LMU Munich, Germany• September 2009 PreDoc school on Ultracold Quantum Gases of Atoms andMolecules, Les-Houches, France• November 2008 Tutor at the ICTP/Democritos Advanced School <strong>in</strong> HighPerformance and GRID Comput<strong>in</strong>g, Trieste, Italy• October 2008 Quo Vadis BEC? conference, Bad Honnef, Germany• September 2007 Path Integrals – New Trends and Perspectives,PI07 Conference, Dresden, Germany• August 2007 V International Student Conference of the Balkan PhysicalUnion, Bodrum, Turkey• July 2007 QTS-5 Conference, Valladolid, Spa<strong>in</strong>• March 2007 ICTP/Democritos Advanced School <strong>in</strong> High PerformanceComput<strong>in</strong>g Tools for e-Science, Trieste, Italy159

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!