• Keine Ergebnisse gefunden

The simplest theoretical model of SPDC assumes the coupling of three discrete modes of field oscillations4. Even though this model is a rough simplification, and a more adequate treatment including multimode description of interacting fields has to be applied (see the next section 2.3), it allows us to readily calculate basic scaling rules of conversion efficiency and deduce some interesting aspects associated with SPDC.

The dynamics of the mode coupling is described by the time-dependent nonlinear interaction Hamiltonian of the form [31]:

HˆI(t) = Z

PiNLEid3r=²0 Z

χ(2)ijkEiEjEkd3r, (2.15) wherePiNL is the nonlinear part of dielectric polarization from Eq. (2.10), and sum-mation on repeated indices is understood. The interacting modes in SPDC must be expressed in the quantized forms in terms of the annihilation ˆa and creation ˆa operators, which obey well-known bosonic commutation relations:

£ˆam,ˆam0

¤ =δmm0, £ ˆ

am,ˆam0¤

=£ ˆ am,ˆam0

¤= 0, (2.16)

where m and m0 are the mode indices. Consequently, the complex-valued electric fields E in (2.15) are transmuted into field operators ˆE. In the most elementary form that takes into account only one possible polarization direction and propagation along the z axis, the electric-field operator is expressed as [32],

E(z, t) =ˆ E0

¡ˆaei(kz−ωt)+ ˆae−i(kz−ωt)¢

, (2.17)

whereE0 is a parameter containing all the prefactors emerging from the field quanti-zation. The first term in relation (2.17) corresponds to the positive-frequency part of the field ˆE(+)(z, t), which is associated with photon absorption, whereas the second corresponds to the negative-frequency part ˆE(−)(z, t), associated with photon emis-sion. Under the assumption of a given phase matching configuration, and considering thatEs in (2.15) formally represent the pump-, signal- and idler-mode operators, the interaction Hamiltonian becomes:

HˆI(t) = 2²0deff Z

−∞

δ(z−z0) ˆEp(+)(z, t) ˆEs(−)(z, t) ˆEi(−)(z, t) dz+ h.c.

= 2²0deffE0pE0sE0i Z

−∞

δ(z−z0)ei∆kzdzˆapˆasaˆie−i(

z }|0 { ωp−ωs−ωi)t

+ h.c., (2.18)

4In principle, this situation could be physically realized by the requirements of phase matching for three modes in a cavity.

where deff is an effective nonlinearity, which can be determined from the tensor d [see definition in the paragraph below Eq. (2.10)] assuming a certain crystallographic structure of the nonlinear medium [30].

The description of the mode coupling using interaction Hamiltonian (2.18) ac-counts for an effect of pump depletion, due to the quantized form of this field. How-ever, under standard experimental conditions this effect is negligible, because the incident pump field is intense and conversion efficiency in SPDC is very low. The latter can be inferred from closer inspection of the relative magnitudes between the linear and the nonlinear term in the expansion of the dielectric polarization:

¯¯PNL¯¯

|PL| = 2deff|Es(i)|

n21 , (2.19)

where |Es(i)| represents the strength of the signal (or idler) mode, which emerges as amplification of vacuum fluctuations; |Es(i)| ¿ 1 V/m. Assuming the realistic magnitudes of the other quantities, deff 10−12 m/V, n2 1 – 10, this ratio is close to zero, corresponding to the spontaneous nature of SPDC. Therefore, to a good approximation, we can treat the pump mode ˆap classically as a complex-valued field of a constant amplitude ap.

The total Hamiltonian consists of the term ˆH0 describing the energy of a free two-mode field and the interaction term ˆHI from Eq. (2.18) [31]:

Hˆ = ˆH0+ ˆHI = X

m=s,i

m µ

ˆ

amˆam+ 1 2

¶ +~g

³ ˆ

aiˆasap+ h.c.

´

, (2.20)

whereg is the mode coupling parameter describing the strength of nonlinear interac-tion. It is proportional to the effective nonlinearity deff and to a factor ei∆ktc, where we putt =z0/c. In Heisenberg representation the time evolution of the field operators is described by the coupled equations of motions [31]:

as dt = 1

i~

h ˆ as,Hˆ

i

=−iωsˆas−igˆaiap, (2.21a) dˆai

dt = 1 i~

h ˆ ai,Hˆ

i

=−iωiˆai −igˆasap, (2.21b) and their Hermitian conjugates. Note that these equations are identical to equations derived for a classical parametric amplifier, see e.g. [29], provided that the annihila-tion and creaannihila-tion operators are identified with classical mode amplitudes and their complex conjugates, respectively. Making use of commutation rules (2.16), it follows directly from Eqs. (2.21):

d

dtˆasˆas = d

dtˆaiˆai, (2.22)

which is equivalent to the commutation relations:

£aˆsˆas,Hˆ¤

=£ ˆ

aiˆai,Hˆ¤

, (2.23)

so that ˆasˆasˆaiˆai is a constant of motion. Recalling the definition of the number operator ˆn [32]:

ˆ

aˆa|ni= ˆn|ni=n|ni, (2.24) wheren is the number of quanta in a mode and |ni is the corresponding eigenstate, we can finally write:

ˆ

ns(t)−nˆs(0) = ˆni(t)−nˆi(0), (2.25) which is a well known Manley-Rowe relation5 [33], reflecting the fact that signal and idler photons are always created in pairs.

The equations of motions (2.21) posses the following solution [34]:

ˆ

as(t) =e−iωst h

ˆ

as(0) cosh¡ κ|ap|¢

−iˆai(0) sinh¡

κ|ap|¢i

, (2.26a)

ˆ

ai(t) =e−iωit h

ˆ

ai(0) cosh¡ κ|ap|¢

−iˆas(0) sinh¡

κ|ap|¢i

, (2.26b)

where we introduced κ(tI) = RtI

−∞g(t)dt. In practice, the interaction time tI may be taken as propagation time through the nonlinear medium of length L, tI L/c, which allows to reduce the integration limits inκ: Rt

−∞−→RtI

0 . The Eqs. (2.26) can be readily used to calculate certain expectations on photon number statistics. To this end we first express the number operators in terms of the field operators att = 0:

ˆ

ns(t) = ˆas(t)ˆas(t) = ˆas(0)ˆas(0) cosh2¡ κ|ap|¢

1 + ˆai(0)ˆai(0)¤

×sinh2¡ κ|ap|¢

1 2i£

ˆ

as(0)ˆai(0)ˆas(0)ˆai(0)¤ sinh¡

2κ|ap|¢

, (2.27a)

ˆ

ni(t) = ˆai(t)ˆai(t) = ˆai(0)ˆai(0) cosh2¡ κ|ap|¢

1 + ˆas(0)ˆas(0)¤

×sinh2¡ κ|ap|¢

1 2i£

ˆ

ai(0)ˆas(0)ˆai(0)ˆas(0)¤ sinh¡

2κ|ap|¢

. (2.27b) Next, assuming that the initial state at t = 0 is |ns(0), ni(0)i, the time evolution of the average photon-numberhnsi (hnsi) at frequencyωsi) can be easily evaluated:

hns(t)i=ns(0) cosh2¡ κ|ap|¢

+ [1 +ni(0)] sinh2¡ κ|ap|¢

, (2.28a)

hni(t)i=ni(0) cosh2¡ κ|ap|¢

+ [1 +ns(0)] sinh2¡ κ|ap|¢

. (2.28b)

5Since the number of photons nis related to the optical power P by P =n~ω, we can rewrite expression (2.25) in the form Pss = Pii, in accordance with the original formulation from Manley and Rowe.

Due to the commutation rules (2.16), the second terms in Eqs. (2.28a) and (2.28b) contain an extra 1, which gives under any initial conditions a nonzero contribution sinh2¡

κ|ap|¢

to the average photon number. Thus, even if the signal and idler modes are initially in vacuum states, i.e. ns(0) = ni(0) = 0, after a time period tI long enough there will be photons in these modes. This purely quantum-mechanical effect elucidates the possibility of spontaneous emission in parametric down-conversion, which emerges as an amplification of the vacuum fluctuations associated with the noncommutation of the field operators. Let us note that the presence of the input signal field stimulates the emission of photons in the idler field and vice versa. That is, the initial conditions ns(0) 6= 0 or ni(0) 6= 0 correspond to the effect of stimu-lated emission, which is fully accounted for by the classical theory of the parametric amplifier.

The interaction time tI is extremely short for realistic crystal lengths (∼mm), so that generally we can consider the short-time limit condition,κ|ap| ¿1, to be valid.

Then, the photon flux emitted from SPDC is given by hns(t)i=hni(t)i= sinh2¡

κ|ap|¢

¡

κ|ap|¢2

. (2.29)

The average photon numbers in the signal and idler mode are proportional to the intensity of the pump fieldIp ∼ |ap|2. AsIpgives the rate at which pump photons fall on the nonlinear medium, the parameter |κ|2 is a dimensionless number determining the fraction of incident pump photons to be converted into lower-frequency photons.

The following scaling behavior can be inferred by closer inspection of the parameter κ, see the definition below Eqs. (2.26):

κ∝deffLsinc

µ∆kL 2

, (2.30)

where sinc function, sinc(x) = sin(x)/x, accounts for the impact of phase mismatch

∆k on the efficiency of SPDC. As illustrated in Fig. 2.2(a), for a given L the phase mismatch ∆k corresponds to a decrease in efficiency by a factor, which is inversely proportional to L. The quadratic scaling of the photon flux with L for the case of perfect phase matching is therefore generally reduced to a linear dependence L if

∆k 6= 0, see Fig. 2.2(b). Furthermore, it follows from (2.30) that the yield of down-conversion photons grows quadratically with the effective nonlinearity deff. Due to the fact that signal and idler photons are always created in pairs, the afore-mentioned scaling rules do not apply only for photon emissions into an individual mode, but also for simultaneous double-photon emissions into both modes.

Notably, the above simple theoretical model is sufficient to prove the nonclassical statistics of down-conversion light [31]. To that end, the mathematical steps leading to Eqs. (2.28) are again applied here to evaluate the second momenth: ˆn2s,i:i and the

Figure 2.2: Effect of phase mismatch on the efficiency of SPDC. (a) The phase mis-match ∆kleads to a decrease of the conversion efficiency in the SPDC process by a factor sinc2(∆kL/2). For ∆k= 0 this factor equals to 1, but with growing ∆kthe factor decreases till it reaches 0 at ∆k = 2π/L. The width of the sinc function is inversely proportional to L, hence clarifying why the condition of phase matching is more restrictive for longer interaction lengths L. (b) The evolution of the average photon number for the case of perfect phase matching, ∆k= 0 and no-phase matching at all, ∆kÀ0, is shown. For real situations including a continuous range of possible values ∆k0, the integration over this range has to be performed to observe the scaling behavior of the photon flux.

cross-correlationh: ˆnsˆni:i6:

h: ˆn2m(t) :i=h0,0|ˆa†2m(t)ˆa2m(t)|0,0i= 2 sinh4¡ κ|ap|¢

, m= s,i, (2.31a) h: ˆns(t)ˆni(t) :i=h0,0|ˆas(t)ˆai(t)ˆai(t)ˆas(t)|0,0i= 2 sinh2¡

κ|ap|¢£

1 + sinh2¡

κ|ap|¢¤

, (2.31b) so that the following inequality must hold :

h: ˆns(t)ˆni(t) :i> 1 2

¡h: ˆn2s(t) :i+h: ˆn2i(t) :i¢

. (2.32)

This is however at variance with the analogical classical inequality:

hI1I2i ≤ 1 2

¡hI12i+hI22i¢

, (2.33)

which evidently has to be fulfilled for any arbitrary classical intensities I1 and I2, because 0≤ h(I1−I2)2i =hI12i+hI22i −2hI1I2i. Ample experimental corroboration of the non-classical character of down-conversion light has been achieved [35, 36].

6The colons denote the normal order of operators, in which all the creation operators stand to the left from all the annihilation operators, such that the vacuum expectation value of the normally ordered product is zero.