You are on page 1of 7

Reflection-mode optical injection locking

Weijian Yang,1 Peng Guo,1,2 Devang Parekh,1 and Connie J. Chang-Hasnain1,*


1

Department of Electrical Engineering and Computer Sciences, University of California, Berkeley, California 94720, USA 2 State Key Laboratory of Advanced Optical Communication Systems and Networks, School of Electronics Engineering and Computer Science, Peking University, Beijing 100871, China *cch@eecs.berkeley.edu

Abstract: A novel model including the interference effect of master laser reflection is established for reflection-mode optical injection locking. This model sheds insight on the physical origin of some rather distinct but unexplained modulation characteristics of optical injection-locked verticalcavity surface-emitting lasers (VCSELs), including data pattern inversion in on-off keying modulation, a large RF gain at low frequency, and an anomalous DC-suppression under small signal modulation, in specific locking conditions. Excellent agreement is obtained between the simulation and experiment results.
2010 Optical Society of America
OCIS codes: (140.3520) Lasers, injection-locked; (250.7260) Vertical cavity surface emitting lasers; (260.3160) Interference.

References and links


1. E. K. Lau, X. Zhao, H. K. Sung, D. Parekh, C. Chang-Hasnain, and M. C. Wu, Strong optical injection-locked semiconductor lasers demonstrating > 100-GHz resonance frequencies and 80-GHz intrinsic bandwidths, Opt. Express 16(9), 66096618 (2008). X. Zhao, B. Zhang, L. Christen, D. Parekh, W. Hofmann, M. C. Amann, F. Koyama, A. E. Willner, and C. J. Chang-Hasnain, Greatly increased fiber transmission distance with an optically injection-locked vertical-cavity surface-emitting laser, Opt. Express 17(16), 1378513791 (2009). D. Parekh, B. Zhang, X. Zhao, Y. Yue, W. Hofmann, M. C. Amann, A. E. Willner, and C. J. Chang-Hasnain, 90-km single-mode fiber transmission of 10-Gb/s multimode VCSELs under optical injection locking, in Optical Fiber Communication Conference, OSA Technical Digest (CD) (Optical Society of America, 2009), paper OTuK7. L. Chrostowski, X. Zhao, and C. J. Chang-Hasnain, Microwave performance of optically injection-locked VCSELs, IEEE Trans. Microw. Theory Tech. 54(2), 788796 (2006). X. Zhao, and C. J. Chang-Hasnain, A new amplifier model for resonance enhancement of optically injectionlocked lasers, IEEE Photon. Technol. Lett. 20(6), 395397 (2008). E. K. Lau, H. Sung, and M. C. Wu, Frequency response enhancement of optical injection-locked lasers, IEEE J. Quantum Electron. 44(1), 9099 (2008). A. Murakami, K. Kawashima, and K. Atsuki, Cavity resonance shift and bandwidth enhancement in semiconductor lasers with strong light injection, IEEE J. Quantum Electron. 39(10), 11961204 (2003). T. B. Simpson, J. M. Liu, and A. Gavrielides, Bandwidth enhancement and broadband noise reduction in injection-locked semiconductor lasers, IEEE Photon. Technol. Lett. 7(7), 709711 (1995). F. Mogensen, H. Olesen, and G. Jacobsen, Locking conditions and stability properties for a semiconductor laser with external light injection, IEEE J. Quantum Electron. 21(7), 784793 (1985). R. Lang, Injection locking properties of a semiconductor laser, IEEE J. Quantum Electron. 18(6), 976983 (1982). W. Yang, P. Guo, D. Parekh, W. Hofmann, M. C. Amann, and C. J. Chang-Hasnain, Physical origin of data pattern inversion in optical injection-locked VCSELs, in Frontiers in Optics, OSA Technical Digest (CD) (Optical Society of America, 2009), paper FTuW2. P. Guo, W. Yang, D. Parekh, W. Hofmann, M. C. Amann, and C. J. Chang-Hasnain, Anomalous modulation characteristics of optical injection-locked VCSELs, in Asia Communications and Photonics Conference and Exhibition, Technical Digest (CD) (Optical Society of America, 2009), paper TuD6. C. H. Henry, N. A. Olsson, and N. K. Dutta, Locking range and stability of injection locked 1.54 m InGaAsP semiconductor lasers, IEEE J. Quantum Electron. 21(8), 11521156 (1985). W. Hofmann, N. H. Zhu, M. Ortsiefer, G. Bhm, Y. Liu, High speed (>11 GHz) modulation of BCB-passivated 1.55 m InGaAlAs-InP VCSELs, Electron. Lett. 42, 976977 (2006).

2.

3.

4. 5. 6. 7. 8. 9. 10. 11.

12.

13. 14.

#130414 - $15.00 USD

Received 22 Jun 2010; revised 1 Sep 2010; accepted 12 Sep 2010; published 12 Oct 2010

(C) 2010 OSA

27 September 2010 / Vol. 18, No. 20 / OPTICS EXPRESS 20887

15. X. Wang, B. Faraji, W. Hofmann, M.-C. Amann, and L. Chrostowski, Interference effects on the frequency response of injection-locked VCSELs, in The 22nd IEEE International Semiconductor Laser Conference (Institute of Electrical and Electronics Engineers, New Jersey, 2010), poster P11.

1. Introduction Optical injection locking (OIL) has been shown to greatly enhance the performance of a directly-modulated laser [18]. Resonance frequency greater than 100 GHz and an intrinsic 3 dB bandwidth up to 80 GHz have been demonstrated for directly modulated vertical-cavity surface-emitting laser (VCSEL) under strong OIL [1]. In this regard, similar behavior is found between injection locking of a VCSEL and an edge-emitting distributed feedback (DFB) laser [1]. Recently, we reported an interesting data pattern inversion and the corresponding chirp polarity inversion in both single transverse mode and multi-transverse mode OIL-VCSELs. This leads to a ten-fold and ninety-fold increased single-mode fiber transmission distance respectively [2,3]. This never-predicted-before phenomenon has only been observed in VCSELs. In addition, there are two other anomalies in small-signal modulation characteristics (S21) of OIL-VCSELs that cannot be explained by standard OIL rate equations [610]. First, a large 20 dB RF gain at low frequency can be obtained for OIL-VCSEL at certain detuning values (wavelength difference between the master and the free-running slave laser) [4]. Second, under certain other detuning values, a dip or zero occurs at very low frequency, herein referred as DC-suppression. In this paper, we establish a novel model [11,12] which includes the interference effect of the master laser reflection for the first time. This interference was completely ignored in all the prior analyses of the modulation characteristic of reflection-mode OIL-lasers. With the inclusion of this interference, the distinct modulation features of OIL-VCSELs can be fully explained and excellent agreements are obtained with experimental results. The model established in this paper will be suitable for all reflection-mode OIL-lasers including VCSELs and some edge-emitting lasers. The reflection not only differentiates the reflection-mode OILlasers from the transmission-mode OIL-lasers, but can be leveraged for further performance improvements, which is also discussed here. 2. Theoretical modeling and simulation Figure 1 shows the schematic of an OIL-VCSEL. The master laser field Em impinges onto the front facet of the slave VCSEL and is divided into transmission and reflection. The transmission interacts with the VCSEL cavity, described by the standard rate equations [6 10]. A steady state is reached inside the cavity and its output Es is phase coherent with Em, with a phase shift s ranging from 0.5 to cot1 [6,9], where is the linewidth enhancement factor. s is determined by the detuning = m - s, where m and s are the wavelength of the master laser and the free-running slave VCSEL respectively. The total output field Etotal is the sum of Es and the reflection of the master laser Emr, where r is the reflectivity of the top facet of the VCSEL, with a phase shift r depending on VCSELs top DBR structure and m. r is around in general. In previous models, only transmitted light was considered and, hence, the interference between Es and Emr was ignored. The total output power Ptotal of the steady state can be written as,

1 1 1 2 2 2 Etotal = Es + Em r 2 + Es Em r cos (s r ) . (1) 2 2 2 Figure 2a shows the output power in the locking map for a transmission-mode OIL-laser [6], while Fig. 2b shows the output power for a reflection-mode OIL-laser where the interference effect of master laser reflection is taken into account. The total output power increases with injection ratio (defined as 20log10(|Em|/|Efr|), where Efr is the electric field of the free-running slave laser light), which is the same in both cases. In the reflection-mode OIL, however, Ptotal decreases with detuning, opposite to the case of the transmission-mode Ptotal =

#130414 - $15.00 USD

Received 22 Jun 2010; revised 1 Sep 2010; accepted 12 Sep 2010; published 12 Oct 2010

(C) 2010 OSA

27 September 2010 / Vol. 18, No. 20 / OPTICS EXPRESS 20888

OIL [6,13]. This is because s increases from 0.5 to cot1, leading to a more destructive interference and thus a lower total output power.

Fig. 1. OIL-VCSEL model with the interference effect. Total output field Etotal = Es + Er.

Fig. 2. The total output power for (a) transmission-mode OIL and (b) reflection-mode OIL.

Small signal analysis is then performed based on the rate equations [6,7,9,10] with the reflection model. Both |Es|2 and phase s have a response under the small-signal modulation, written as |Es|2 and s respectively, which are superimposed on the steady state solution. Based on Eq. (1), the total output power is written as, 1 2 2 Etotal + Etotal 2 (2) 1 1 2 2 2 2 2 1 = Es + Es + Em r 2 + Es + Es Em r cos (s + s r ) . 2 2 2 With the main simulation parameters listed in Table 1 in Appendix, the typical frequency response of Ptotal is simulated for different detuning values under a strong injection ratio. The results are shown in Fig. 3, where the amplitude response |Ptotal|2 (the square converts optical domain into electrical domain, known as S21) is normalized to the free-running case. It is clearly seen that there is a DC-suppression in the amplitude response as the detuning increases from blue (<0) to red (>0). This DC-suppression corresponds to a phase change in the phase response (phase of Ptotal). It is the destructive interference between the OIL-VCSEL internal output field Es and the master laser reflection light Emr that leads to this DCsuppression and phase change. We will show later that this corresponds to the transition point for data pattern inversion in large signal modulation. As the detuning value increases further, the DC-dip disappears and a very large RF gain is obtained, again due to the interference effect. Ptotal + Ptotal =

#130414 - $15.00 USD

Received 22 Jun 2010; revised 1 Sep 2010; accepted 12 Sep 2010; published 12 Oct 2010

(C) 2010 OSA

27 September 2010 / Vol. 18, No. 20 / OPTICS EXPRESS 20889

Fig. 3. Simulation results for OIL-VCSEL small-signal frequency response (S21) with different detuning values under a strong injection ratio of 16.15 dB. (a) Normalized amplitude response; (b) Phase response. The arrow on (b) indicates a phase change as the detuning increases from blue to red. From blue detuning to red detuning, the detuning values for the curves are 0.602 nm, 0.199 nm, 0.822 nm, 1.500 nm, 1.883 nm; corresponding to the phase shift s 0.490, 0.4486, 0.400, 0.300, 0.165, respectively.

Data pattern under on-off keying (OOK) large-signal modulation is simulated by fourthorder Runge-Kutta method. Extinction ratio re is defined as 10log10(P1/P2), where P1 and P2 are the OIL-VCSEL output powers corresponding to the high and low level of the modulation current. Thus a negative extinction ratio indicates data pattern inversion. Figure 4 shows the typical 1 Gb/s data patterns for different detuning values, with a fixed injection ratio. The average total output power decreases with detuning, as predicted by Fig. 2b. With increasing detuning, the data pattern changes from normal to transition state, and then to inverted.

Fig. 4. Simulation results for OIL-VCSEL 1 Gb/s OOK data pattern with different detuning values under a fixed injection ratio of 12.0 dB. As the detuning changes from blue to red, the data pattern changes from normal to transition state, and then to inverted. From blue detuning to red detuning, the detuning values for the curves are 0.115 nm, 0.321 nm, 1.178 nm, 1.662 nm; corresponding to the phase shift s 0.470, 0.430, 0.300, 0.125, respectively.

Next, we compare simulation results for small- and large- signal modulation response by sweeping the parameter space of injection ratio and detuning value. Figure 5 shows the RF response at 1 GHz and the extinction ratio of the 1 Gb/s data pattern on the same locking map. The locking range of the large-signal modulation is slightly smaller than that of the smallsignal modulation, due to the larger perturbation to the system in the large-signal modulation. It is interesting to note that the DC-suppression in Fig. 5a is at the same parameter space with the transition state at which the extinction ratio is zero in Fig. 5b, confirming a strong correlation between them. For a certain injection ratio at the red detuning side, the interference effect makes P1 into a very small value, resulting in a very large (in magnitude), and desirable extinction ratio with data pattern inverted. This is indeed the regime of interests for greatly increased fiber transmission distance [2,3].

#130414 - $15.00 USD

Received 22 Jun 2010; revised 1 Sep 2010; accepted 12 Sep 2010; published 12 Oct 2010

(C) 2010 OSA

27 September 2010 / Vol. 18, No. 20 / OPTICS EXPRESS 20890

Fig. 5. Simulation results for the RF response of the small-signal analysis at 1 GHz and the extinction ratio re of the 1 Gb/s OOK large-signal modulation of OIL-VCSEL on the same locking map. The negative re indicates data pattern inversion. The dashed line on (b) indicates the conditions where re = 0.

3. Experiment

Detailed experiments on 1.55 m buried-tunnel-junction (BTJ) VCSELs [14] were performed to verify the simulation results and, in particular, the correlation between small- and largesignal modulation. The VCSEL and experimental setup were similar to that reported elsewhere [1]. The VCSEL was biased at 4 mA, 2.5 times threshold current, with 3 dBm free-running output power. The small-signal modulation response of the OIL-VCSEL was measured at a fixed injection ratio of 20 dB for different detuning values, shown in Fig. 6. At blue detuning (e.g. = 0.759 nm), a high resonance frequency of 100 GHz was attained, similar to what was reported [1]. The large damping at lower frequencies is due to the parasitic RC of the VCSEL. Increasing the detuning to 0.149 nm, a large DC-dip was seen, which was not reported before. Further detuning to the red side, the DC-dip disappeared and a large RF gain 12.5 dB at low frequency was obtained. Similar observation with >20 dB gain was reported in Ref. 4.

Fig. 6. RF response of the small-signal modulation of the OIL-VCSEL for different detuning values (0.759 nm, 0.149 nm, 0.395 nm and 1.165 nm), at a fixed injection ratio of 20 dB. FR: free running.

Figure 7 shows experimental results of OIL-VCSEL under 1 Gb/s OOK modulation at different detuning values, with the injection ratio optimized at 12.9 dB for good extinction ratios. With increasing detuning, the data pattern changes from normal, transition state, to inversion. The typical top DBR mirror field reflectivity of the VCSEL is higher than 99%, resulting in a strong reflection and interference effect of the master laser light. Extinction ratio as high as 12 dB with a 3.6 dB signal amplification compared to the free-running case was achieved with the 1.036 nm detuning.

#130414 - $15.00 USD

Received 22 Jun 2010; revised 1 Sep 2010; accepted 12 Sep 2010; published 12 Oct 2010

(C) 2010 OSA

27 September 2010 / Vol. 18, No. 20 / OPTICS EXPRESS 20891

Fig. 7. Data pattern of the 1Gb/s OOK large-signal modulation of OIL-VCSEL for different detuning values (0.230 nm, 0.134 nm, 0.630 nm and 1.036 nm), at a fixed injection ratio of 12.9 dB. FR: free running.

The RF response at 1 GHz and the extinction ratio of the 1 Gb/s data pattern were measured at different points in the same parameter space for another VCSEL (5 mA bias current which is 3 times threshold current, 0.7 dBm free-running output power), with the results shown in Fig. 8. The strong correlation between the DC-suppression in small-signal modulation and transition state in large-signal modulation is clearly seen. All these results agree very well with the simulation.

Fig. 8. RF response of the small-signal modulation at 1 GHz and extinction ratio re of the 1 Gb/s OOK large-signal modulation of OIL-VCSEL on the same locking map. The crosses show in which conditions the data were taken in the experiment, and the dashed line on (b) indicates the conditions where re = 0.

4. Application

The reflection-mode OIL model would find its great application in optical communication and optical data processing. It predicts the conditions of data pattern inversion and the corresponding chirp polarity inversion, which is exactly the region where the greatly increased single-mode fiber transmission distance was achieved in the OIL-VCSELs [2,3]. Furthermore, since the OIL-VCSEL can operate either in normal data state or inverted data state, it is possible to develop some optical switching applications. For example, a frequency-modulated master laser can be used to modulate the detuning value of the OIL-VCSEL, and thus the switching can occur between the normal data state and inverted date state. Alternatively, an intensity-modulated master laser can be used to control the injection power and thus the switching. However, as seen in Fig. 5 and Fig. 8, the transition line is very flat with respect to the injection ratio, making this switching difficult. On the other hand, by adjusting the VCSELs front DBR reflectivity and phase, it is possible to tilt the transition line. Figure 9 shows the same simulations as Fig. 5, except that the front DBRs field reflectivity is changed from 0.9968 to 0.9962 while the reflection phase from 1.000 to 0.970 at 1550 nm. The

#130414 - $15.00 USD

Received 22 Jun 2010; revised 1 Sep 2010; accepted 12 Sep 2010; published 12 Oct 2010

(C) 2010 OSA

27 September 2010 / Vol. 18, No. 20 / OPTICS EXPRESS 20892

transition line is tilted, and the output of the VCSEL can switch between normal state and inverted state by different injection powers from the master laser. Reference [11] shows the initial experimental switching results, and this may be possibly further developed into an optical logic gate.

Fig. 9. Simulation results for the extinction ratio re of the 1 Gb/s OOK large-signal modulation of the OIL-VCSEL on the locking map. (Simulation parameters are listed in Table 1 in Appendix, except that the field reflectivity of the VCSELs front DBR is changed from 0.9968 to 0.9962 while the reflection phase from 1.000 to 0.970 at 1550 nm.) The dashed line indicates the conditions where re = 0. The output of the VCSEL can switch between normal state and inverted state by different injection powers from the master laser, shown by the double head arrow line.

5. Conclusion

The proposed model incorporates the interference effect of the master laser reflection into the total output of the reflection-mode OIL-lasers. This reflection, which once was considered detrimental for practical applications, is shown to be a key to explain the rather distinct modulation characteristics of OIL-VCSELs. The small-signal RF response is strongly correlated to the data patterns in large-signal modulation. The distinct DC-suppression reflects itself into the transition state in large-signal modulation. The large RF gain at low frequency is both seen in simulation and experiment. This model will not only be useful to optimize the conditions for OIL-VCSEL-based optical communication, but also expand OIL-VCSELs applications into multi-level modulation, photonic switching and logic gates.
Appendix

Simulation parameters of the slave VCSEL are listed in the following table.
Table 1. Simulation Parameters of the Slave VCSEL Central wavelength Top DBR field reflectivity Bottom mirror field reflectivity Cavity length Differential gain Transparency carrier density Cavity distributed material loss 1.55 m 0.9968 1 2.87 m 5.5 1020 m2 2.0 1024 m3 3.8 103 m1 Mode confinement factor Linewidth enhancement factor Photon life time Carrier life time Gain saturation coefficient Laser active region volume Cavity effective refractive index 0.048 8 2.2 1012 s 8.3 1010 s 2.7 1023 m3 1.0 1018 m3 3.24

Acknowledgements

The authors thank M. C. Amann and W. Hofmann for providing the 1.55 m VCSELs, and the support from National Security Science and Engineering Faculty Fellowship. The authors were made aware of an independent work [15] with relevant similarities after the submission of this manuscript.

#130414 - $15.00 USD

Received 22 Jun 2010; revised 1 Sep 2010; accepted 12 Sep 2010; published 12 Oct 2010

(C) 2010 OSA

27 September 2010 / Vol. 18, No. 20 / OPTICS EXPRESS 20893

You might also like