Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Roadmap for gain-bandwidth-product enhanced photodetectors: opinion

Open Access Open Access

Abstract

Photodetectors are key optoelectronic building blocks performing the essential optical-to-electrical signal conversion, and unlike solar cells, operate at a specific wavelength and at high signal or sensory speeds. Towards achieving high detector performance, device physics, however, places a fundamental limit of the achievable detector sensitivity, such as responsivity and gain, when simultaneously aimed to increasing the detector’s temporal response (speed) known as the gain-bandwidth product (GBP). While detector’s GBP has been increasing in recent years, the average GBP is still relatively modest (∼106-109 Hz-A/W). Here we discuss photoconductor-based detector performance limits and opportunities based on arguments from scaling length theory relating photocarrier channel length, mobility, electrical resistance with optical waveguide mode constrains. We show that short-channel detectors are synergistic with slot-waveguide approaches, and when combined, offer a high-degree of detector design synergy especially for the class of nanometer-thin materials. Indeed, we find that two-dimensional material-based detectors are neither limited by their low mobility nor by associated carrier velocity saturation limitations and can, in principle, allow for 100 GHz fast response rates, which is unlike traditional detector designs that are based on wide channel lengths. However, the contact resistance is still a challenge for such thin photo absorbing materials – a research topic that is still not addressed yet. An interim solution is to utilize heterojunction approaches for functionality separation. Nonetheless, atomistic and nanometer-thin materials used in such next-generation scaling length theory based detectors also demand high material quality and monolithic integration strategies into photonic circuits including foundry-near processes. As it stands, this letter aims to guide the community if achieving the next generation photodetectors aiming for a performance target of GBP ∼ 1012 Hz-A/W.

© 2020 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

Corrections

9 September 2020: A typographical correction was made to the article title.

1. Main text

Photodetectors deliver an integral function of optical-to-electrical (OE) conversion. Unlike solar cells performing this conversion spectrally broadband and temporally at steady-state, photodetectors, and especially photoreceivers are typically designed for high responsivity (gain) at a target wavelength given by the absorbing materials’ bandgap, and by a fast-temporal dynamic response (3 dB bandwidth, BW). Prominent material-spectrum mappings for detectors are GaN for UV (<400 nm) [1], Si for visible to NIR (400-1100 nm) [2], Ge/InGaAs for NIR to MIR (1-5 µm) [3] and HgCdTe for MIR to FIR (>5 µm) [4].

It is the aim of this letter to clarify i) fundamental photodetector tradeoffs, and ii) relate them to design choices, which thence govern detector performance. We show that this correlation is actually not arbitrary and that there exist interdependencies between design-material combinations that yield higher performance than others. The arguments laid out herein, are exemplary made on photonic integrated circuit (PIC)-based detectors, but their generality also holds for free-space coupled devices and applications.

The obtainable signal output from a detector (i.e., photocurrent and available photogain) and its operation response (speed, i.e. bandwidth, BW) are not interdependently optimizable, which can straightforwardly be understood by using the picture of the integration time of a photo generated signal; if the gain is high to increase the detector's sensitivity, then the response rate is slow, and vice versa. An example from daily life is the sluggish smart-phone camera response for taking photos at dim lighting conditions, where the digital signal processing circuit (DSP) increases the gain to enhance the signal-to-noise ratio (SNR), thus reducing the shutter speed.

Most of the photodetectors can be categorized broadly into two classes (Fig. 1(a)); those with high responsivity (R) (but low temporal bandwidth, speed) and vice versa, which relates to the fundamental trade-off gain-bandwidth product (GBP) being a figure-of-merit (FOM) for detectors. Since photodetector gain linearly scales with the responsivity in units of (A/W), here, for discussion purposes, we use gain and responsivity interchangeably. Indeed, the iso-GBP lines show devices cluster in either the upper left (low-BW/high-R) or lower right (high-BW/low-R) quadrant, both corresponding to a GBP ∼ 106 - 109 Hz-A/W (Fig. 1(a)). Naturally, detector performance and next-generation devices scale orthogonally to these two quadrants aiming for simultaneous high-speed and high R. It is the target to design and demonstrate such high-performance photodetectors that guide our motivation for the work presented herein (Fig. 1(a)).

 figure: Fig. 1.

Fig. 1. Design strategies for Engineering Gain-Bandwidth-Product (GBP) Photodetectors. a) Photodetector performance cluster into two quadrants; high responsivity yet low speed (bandwidth, BW), and vice versa. Target detectors with a GBP of ∼1012 - 1013 scale orthogonally (upper-right quadrant) to the iso-GBP lines. The values are taken from [526]. b) Plot of bandwidth vs. photocarrier-collecting electrical channel length showing two regimes where the transit time-limited bandwidth (BWtransit) is dominant for long-channel detectors (gray shade), whereas for sub 100 nm channel lengths, the bandwidth is limited by RC time (blue shade). BWtransit is estimated for four different mobility values (1 - 1000 cm2/Vs) for a Vsd = 1 V taken from Ref. [19, 2729]. The RC time-limited bandwidth (BWRC) is estimated for four different resistance values starting from 0.1−100 kΩ [3032], where, the electrical capacitance is determined by a parallel-plate model (fringe fields ignored) for a lateral junction. A finding of this parametric study is that for a (arbitrarily selected) target speed of 100 GHz; i) micrometer long channel-based photodetectors required a rather high mobility (<104 cm2/Vs) but with a somewhat relaxed contact resistance RC requirements, yet high mobilities challenges achievable resistance due to scattering events induced by higher doping, while ii) 10’s nm short-channel detectors are only resistance limited, since (near) ballistic transport is given even for poor mobilities (<10 cm2/Vs), and relatively high grain can be achieved by reducing the transit time to ∼ps as supposed to increasing the carrier lifetime. Such short-channel performance-detectors, however, demand optical mode squeezing to adhere to device scale-length theory (SLT) rules which can be achieved with slot-waveguide designs.

Download Full Size | PDF

Let us next analyze this fundamental GBP tradeoff in more detail to explore device paradigms that would allow to optimize the GBP beyond what is available with current technology; the photoconducting gain is given by the ratio of the photocarrier lifetime (τc) to the photocarrier transit time (τt), i.e. the drift time for photocarriers to reach the photocurrent-collecting electrodes. Meaning, if the carrier lifetime exceeds the time for their collection, charge accumulation (gain $ \approx $ τct) can occur. However, this well-known ‘gain theory’ is highly based on Ohmic contact and on the assumption that the concentration of photogenerated excess carriers in the photoconductor is uniformly distributed which is not true for any photodiode (p-n junction or Schottky junction), where the distribution of photogenerated excess carriers is always non-uniform and therefore the photogenerated excess charge carriers readily skewed by the electric field, resulting in the voltage-dependent excess carrier concentration. This can be seen clearly from the continuity equation, and the total photocurrent will not only be due to the drift current, but also have contributions from the diffusion current. Given that the carrier lifetime in typically used detector semiconductors is about O(∼ns), gains of about O(∼102 - 103) can be obtained for transit times of 10’s - 100’s of ps corresponding to medium-high mobility materials and micrometer source-drain distances. The transit time, on the other hand, is a function of the electrical transport properties (mobility, μ) of the photo absorbing material and the drift field (Vbias) via ${\tau _t} = {L_{e/h}}v_{drift}^{ - 1}$, where Le/h where is the distance of the electrodes collecting the electron-hole photocarriers (i.e. channel length, Le/h).

Parametric BW-scaling of this mobility-limited transit time and RC-delay as a function of photodetector source-drain channel length, Le/h, reveals several insights (Fig. 1(b)); a) If the channel length is several micrometer long, mobilities of about 1,000 cm2/Vs are required to even achieve 1 GHz fast detection, while RC-delay is not a factor even for low resistivities. b) In the limit for scaled (1 - 10 nm) short channel lengths even a poor mobility of 1 cm2/Vs is not BW limiting up to about 100 GHz, provided at least a resistance of 1 kΩ can be guaranteed. c) A sweet spot for detectors that are aimed for speeds around 10’s of GHz is found for channel lengths near 0.1 µm offering a high design window with relatively relaxed requirements (µ < 10 cm2/Vs and R < 100 kΩ). As an interim conclusion, we can summarize that if high-speed devices (>100 GHz) are desired, then short transit channel lengths (∼10’s of nm) are a possible design option for materials with medium to low mobilities. Interestingly, even high-mobility (∼10,000 cm2/Vs) materials cannot reach such high BW for the case where micrometer wide contacts are used, which is the case for many low-index material waveguides such those based on SiO2 and also often Lithium-niobate (LN). Furthermore, we see (Fig. 1(b)) that for III-V and silicon-based waveguides a channel length of 500 nm is already challenging, if plasmonic losses from too-closely placed electrical contacts are to be avoided. Thus, for our exemplary 100 GHz goal and micrometer-far apart contacts, mobilities of ∼1,000 cm2/Vs are required. The latter, however, limits the design window for achievable low-resistance (i.e. highly-doped contacts) due to increases scattering events at high doping levels, which increases resistance. Thus, an interesting design option seems to emerge, which is to accept possible optical losses from metal-optics and design 10’s nm narrow slot-waveguides, which can achieve 100 GHz devices with both modest mobilities ∼5 cm2/Vs and resistances ∼5 kΩ. Such material is, for instance, the class of transition metal dichalcogenides (TMDC). In fact, one can show that borrowing device scaling laws from electronics, that ‘flatness’ of the detector material is a key requirement, as discussed further below.

However, since the gain is proportional to both the mobility and the carrier lifetime, the former may be limiting the bandwidth of the detector such as for poor mobilities known form TMDCs [19]. Yet, there are several reports on ultrahigh gain in TMDCs based photodetector where the underlying mechanism is mainly photogating effect instead of multicarrier generation or avalanche amplification [8,9]. In the photogating effect, the charge carriers are trapped in surface impurity states, thus resulting in a DC-slow response time (1 - 10 s), which is too slow for many data communication and signal processing/computing applications, but possibly sufficient for environmental sensing applications [5,6,8,9,24]. Therefore, for long-channel detectors, in order to achieve high GBP, focus should be on improving carrier mobility rather than prolonging the carrier-lifetime to transit-time ratio, while the inverse is required (i.e. focus on improving contact resistance) for short-channel detectors, as further discussed next (Fig. 1(b)).

Borrowing device scaling concepts from electronics, the scale-length theory (SLT) shows that electric field-based device dimensions scale with a nominal factor (σ) such as in FETs both the channel thickness and length [33]. Applying this dimensional scaling concept to the GBP of photodetectors, we find that it scales as GBP ∼ σ–2Le/h−2, which suggests that reducing the collection distance rather sensitively (squared) improves detector performance. Hence the question arises, as of why prevailing integrated detector designs do not make use of this scaling performance gains?

The answer can be found in the inability of monolithic material integration to adhere to the SLT scaling; starting our GBP optimization analysis discussion with optical constrains, utilizing this σ−2 scaling law urges device designers to utilize nanometer (or 10’s of nanometer) short-channel (contact distances) to optimized GBP. This, however, requires a simultaneous ‘squeezing’ of the optical mode to the same order (10’s nm compact), which can be realized with nano-optical light-matter-interaction enhancements such as offered by plasmonics [34] or dielectric discontinuities [35], typically at the cost of a non-zero metal-loss component, however. SLT applied to detectors simultaneously demands nanometer-thin absorbing materials, which is further enforced to ensure a high optical mode overlap (Γ) [36], similar to quantum-well lasers [37]. Realizing such a nanometer-thin absorbing material at high crystalline quality, however, is challenging to be provided by III-VI and IV (e.g. Ge) materials when these materials are heterogeneously integrated in Si or SiN photonic platforms due to the parasitic high defect density during patterning (e.g., reactive ion etching) [38,39]. Epitaxial direct growth of heterojunctions would be the ideal scenario; however, lattice mismatches severely limit this option, along with thermal budgets and possibly low manufacturing throughput. However, a SLT-scaled device detector is synergistic with two-dimensional (2D) materials, as briefly mentioned above and further discussed next, where an interesting design concept emerges that is based on reducing transit time as supposed to increasing carrier lifetime (Fig. 1).

As a possible high GBP-design example, a rather short plasmonic slot atop a silicon on insulator (SOI) waveguide, where Le/h = Wslot (slot-width ∼10 nm), supports a gap plasmon with an electric field polarization being in-plane with such a nanometer thin material suspending this gap (see Ref. [22,4041] for design examples). In general, the detector’s external quantum efficiency is given by: EQE = IQE x (1-Reflection-Transmission), where the internal quantum efficiency (IQE) is proportional to the overlap matrix element in the absorption process (typically ∼80’s - 90’s% for band-to-band absorption [42]). The second term relates to the amount of light absorbed by the device, which can be optimized by anti-reflection coatings for free-space coupled devices. In integrated devices, however, this relates to the mode overlap factor (Γ), i.e. the amount of light coupled to the photocarrier-generating material(s). Since a detector’s photo responsivity is proportional to the EQE, and hence to GBP, naturally ensuring a high optical absorption is critical. With 2D materials featuring a rather high intrinsic absorption coefficient (∼105 cm−1), they are an interesting SLT-synergistic material option; furthermore, their atomistically-thin flatness reduces Coulomb scattering effects, thus allowing for a high oscillator strength [43] signified by a high optical index Re{n} > 5 in the visible and NIR bands for TMDCs [44]. Furthermore, such a ∼10 nm short electrical channel, also comes with (near) ballistic electric transport; hence, a high drift mobility would actually not improve the transit time which is ∼1 ps, for Vbias = 1 V and μTMDC = 1-10 cm2/Vs, and the low mobility of 2D materials (aside Graphene) is not a limit of an SLT-scaled high-GBP 2D material-based detector designs.

However, improving a material’s mobility does relax length-scaling requirements (const. = GBP → µ ∼ (Le/h)2 (Fig. 1(b)). For instance, some recent undoped TMDC’s have shown mobilities of 103 cm2/Vs [27,28], which allows increasing Le/h from 10’s to 100’s nm for high-speed detectors. Channel-length scaling, however, also impacts the detector’s dark current (Idark); generally speaking, lowering Idark increases the detectors dynamic range, and improves the noise-equivalent-power (NEP), which is the required optical power to produce an SNR of unity at 1 Hz signal rate. Indeed, we [19] and others [20] have shown, that a TMDCs-based detector is able to operate at 1-10 nA dark currents, while a scaled slot detector has an about 10x higher Idark. Compared to graphene, TMDCs feature a 10 - 1000x lower dark current due to their 0.5 - 2 eV wide bandgap, whilst even the smallest energy (optical or thermal) will initiate intra-band transitions in a gap-less material such as graphene. Another possibility for reducing the dark current, is to operate the detector in photovoltaic mode, namely at zero bias voltage, instead of the often used 1 - 2 V, which, however, also gives up gain enhancement design options that are based on carrier lifetime-to-transit time ratios. Nonetheless, such a detector can be realized with built-in P/N junction for charge separation polarity; for instance, surveying the literature, points to a potential of 10 - 100x lower Idark [23,24].

However, 2D materials also have challenges in designing detector (e.g. high resistance), however as discussed before, their low-mobility is not one of them, provided SLT short Le/h approaches are used. Yet, total resistance of the device comprises of contact resistance (Rc) and channel resistance (Rch), which is key to achieving high-performance; since Rch quickly decreases with the channel length scaling, Rc becomes a dominant factor in the total resistance for sub-100 nm channel lengths. For instance, the high contact resistance in kΩ to even MΩ challenges the device response time limiting RC-delay [22] (Fig. 1(b)). A possible solution for this is to design hetero-junctions, where function-material separation can be part of the design process; for instance, a high bandgap TMDC placed atop an SOI waveguide performs the OE conversion, whilst the photocarriers are transferred into graphene layers acting as an electrical conducting channel or as a ∼50 Ω contact resistance to the metal contact pads [30]. Several contact engineering techniques including edge contact geometry, 1T phase contacts, hBN substrate and graphene have been implemented in individual devices, but a scalable route for uniformly realizing these device concepts at the wafer-scale has not yet to be established [45,18].

Taking the above discussed high-GBP design options into account on high gain and responsivity via short transit-times, heterojunction devices for low RC, we can estimate a speed of up to ∼ 100 GHz (limited by RC, not by transit time) and a responsivity of 10 - 100 A/W with a typical TMDC carrier lifetime ∼O(100’s ps) [46], the transit time is only ∼1 ps given a slot width of, exemplary, Le/h = 33 nm and a mobility of 10 cm2/Vs, (assuming elsewise ideal conditions), thus enabling a GBP approaching 1012 - 1013 Hz-A/W (blue region, Fig. 1(a)). This would be about 4 - 5 orders improvement compared to the average detector performance and 1 - 2 orders enhancement over record demonstrations (Fig. 1(a)) [2124]. The resulting Idark would be about O(100’s pA) for a NEP of ∼1 - 10 pW/Hz0.5 in line with state-of-the-art detectors [11].

Interestingly, 2D materials are being actively investigated to reaching foundry maturity such as pursuit by IMEC [47] or TSMC [48], for example, yet no customer-ready process is available as of this point in time of writing this article. This momentum bears hope for further material improvements with R&D economy-of-scale testing in the near future, and monolithic growth on SOI-oxide surfaces is possible, yet quality levels such as defect density still required attention. While CVD such processes are actively being developed to date, transfer techniques for 2D materials have successfully addressed issues such as deterministic placement [49] and single-flake transfer showing vanishing cross-contamination past transfer [50]. For instance, micro-stamping techniques allow for rapid-prototyping devices benign to PIC heterogeneous 2D material integration enabled by the high spatial selectivity of the micro-stamper [51].

Furthermore, 2D materials offer a higher degree of strain-material adaptation given added range of design flexibility, which we termed ‘strainoptronics’; for instance, straining (tensile or compressive) dramatically (100’s of meV) shifts the band edge of 2D materials leading to enhanced absorption so as responsivity for 2D materials based photodetector [19]. The understanding of the modulation of optical and electronic properties of these materials is key to design improved integrated photonic devices [5254]. Techniques such as Kelvin probe force microscopy (KPFM) allow for local strain mapping which allows for nanometer-resolution strain and hence bandgap maps, thus allowing to align the photoabsorbing 2D material with photonic structures such as waveguides, or optical cavities precisely [19]. Indeed, the large amount of strain that 2D materials can sustain before rupture occurs along with their large elastic modulus [55], which allows reducing the bandgap of, e.g. MoTe2, from 1.04 eV down to 0.80 eV, together enable efficient photo conversion at the telecommunication relevant wavelengths of 1550 nm [19].

In conclusion, similar to other opto-electronic fundamental building blocks such as lasers and modulators, the device physics of photodetectors dictate fundamental trade-offs spanning an orthogonal performance space; for a detector, this means that the sensitivity such as the available gain, the responsivity and hence the noise-equivalent power cannot be independently optimized without trade-offs in bandwidth, speed, as both are intercorrelated. Here we discussed photodetector gain-bandwidth-product (GBP) performance in the context of device channel scaling, electrical and opto-electronic performance constrains. We show that state-of-the-art detectors are either optimized for gain or for speed showing an averaged GBP of 106 to 109 Hz-A/W (using responsivity for gain due to linear proportionality amongst them). Further, we discuss strategies for achieving next-generation GBPs of approaching 1012 to 1013 Hz-A/W following scaling length theory, which would constitute an enhancement of one to two orders of magnitude over current best demonstrations. For instance, we show that for achieving 100 GHz-fast detectors, designs using micrometer-wide separated photocarrier-collecting source/drain contacts require exceedingly-high mobilities (∼104 cm2/Vs), which challenges simultaneously achieving low contact resistances in order to ensure short RC-delay. A seemingly more elegant and PIC-density benign approach of achieving high GBP detectors, may be to scale the channel length into the 10’s nanometer-short regime, where transport is near ballistic even for low-mobility (<10 cm2/Vs) materials. Interestingly, we find that the scaling length theory and plasmonic slot-waveguide approaches offer a high-degree of detector design synergies especially for ‘flat’ photo-conversion materials. Such 2D materials may be a solution for such next generation high-GBP devices based on channel length scaling because their low mobility does, in fact, not limit performance, since carrier velocity saturation effects play a minor role and the device response is typically not transit time-limited but determined by RC-delay. However, for such scaled photodetectors, several issues are a limiting factor, foremost of which is the low contact resistance to the channel materials, since capacity increases with scaling, which has to be offset by resistance. The latter is a particular challenge for 2D materials, and future research should place a focus on this. Other challenges, in particular in a context of chip integration, are general scalability and associated yield issues, however with possibilities for tailored doping levels and lack of standardization protocols. Therefore, in order to realize the fullest potential of such device physics-benign emerging materials, constrains from foundry-based mass productions and chip integration should be co-developed to ensure satisfactory and reliable performance.

Funding

Air Force Office of Scientific Research (FA9550-17-1-0377); Army Research Office (W911NF-16-2-0194).

Disclosures

The authors declare that there are no conflicts of interest related to this article.

References

1. Q. Chen, J. W. Yang, A. Osinsky, S. Gangopadhyay, B. Lim, M. Z. Anwar, and M. Asif Khan, “Schottky barrier detectors on GaN for visible–blind ultraviolet detection,” Appl. Phys. Lett. 70(17), 2277–2279 (1997). [CrossRef]  

2. M. K. Lee, C. H. Chu, Y. H. Wang, and S. M. Sze, “1.55-µm and infrared-band photoresponsivity of a Schottky barrier porous silicon photodetector,” Opt. Lett. 26(3), 160–162 (2001). [CrossRef]  

3. G. H. Olsen, A. M. Joshi, and V. S. Ban, “Current status of InGaAs detector arrays for 1–3 µm,” Proc. SPIE 1540, 596–605 (1991). [CrossRef]  

4. A. Rogalski, “HgCdTe infrared detector material: history, status and outlook,” Rep. Prog. Phys. 68(10), 2267–2336 (2005). [CrossRef]  

5. W. Zhang, M. Chiu, C. Chen, W. Chen, L. Li, and A. T. S. Wee, “Role of Metal Contacts in High-Performance Phototransistors Based on WSe2 Monolayers,” ACS Nano 8(8), 8653–8661 (2014). [CrossRef]  

6. J. O. Island, S. I. Blanter, M. Buscema, H. S. J. van der Zant, and A. Castellanos-Gomez, “Gate Controlled Photocurrent Generation Mechanisms in High-Gain In2Se3 Phototransistors,” Nano Lett. 15(12), 7853–7858 (2015). [CrossRef]  

7. N. Huo, S. Yang, Z. Wei, S. Li, J. Xia, and J. Li, “Photoresponsive and Gas Sensing Field-Effect Transistors based on Multilayer WS2 Nanoflakes,” Sci. Rep. 4(1), 5209 (2015). [CrossRef]  

8. O. Lopez-Sanchez, D. Lembke, M. Kayci, A. Radenovic, and A. Kis, “Ultrasensitive photodetectors based on monolayer MoS2,” Nat. Nanotechnol. 8(7), 497–501 (2013). [CrossRef]  

9. J. F. Gonzalez Marin, D. Unuchek, K. Watanabe, T. Taniguchi, and A. Kis, “MoS2 photodetectors integrated with photonic circuits,” npj 2D Mater. Appl. 3(1), 14 (2019). [CrossRef]  

10. N. Youngblood, C. Chen, S. J. Koester, and M. Li, “Waveguide-integrated black phosphorus photodetector with high responsivity and low dark current,” Nat. Photonics 9(4), 247–252 (2015). [CrossRef]  

11. https://www.thorlabs.com Thor Labs, product DETO25A

12. Y. Salamin, P. Ma, B. Baeuerle, A. Emboras, Y. Fedoryshyn, W. Heni, B. Cheng, A. Josten, and J. Leuthold, “100 GHz Plasmonic Photodetector,” ACS Photonics 5(8), 3291–3297 (2018). [CrossRef]  

13. Y. Chen, S. Williamson, and T. Brock, “375-GHz bandwidth photoconductive detector,” Appl. Phys. Lett. 59(16), 1984–1986 (1991). [CrossRef]  

14. F. N. Xia, T. Mueller, Y. M. Lin, A. Valdes-Garcia, and P. Avouris, “Ultrafast graphene photodetector,” Nat. Nanotechnol. 4(12), 839–843 (2009). [CrossRef]  

15. T. Mueller, F. N. A. Xia, and P. Avouris, “Graphene photodetectors for high-speed optical communications,” Nat. Photonics 4(5), 297–301 (2010). [CrossRef]  

16. X. Gan, R. J. Shiue, and Y. Gao, “Chip-integrated ultrafast graphene photodetector with high responsivity,” Nat. Photonics 7(11), 883–887 (2013). [CrossRef]  

17. S. Schuler, D. Schall, D. Neumaier, L. Dobusch, O. Bethge, B. Schwarz, M. Krall, and T. Mueller, “Controlled Generation of a p–n Junction in a Waveguide Integrated Graphene Photodetector,” Nano Lett. 16(11), 7107–7112 (2016). [CrossRef]  

18. R. J. Shiue, Y. Gao, Y. Wang, C. Peng, A. D. Robertson, D. K. Efetov, S. Assefa, F. H. L. Koppens, J. Hone, and D. Englund, “High-Responsivity Graphene–Boron Nitride Photodetector and Autocorrelator in a Silicon Photonic Integrated Circuit,” Nano Lett. 15(11), 7288–7293 (2015). [CrossRef]  

19. R. Maiti, C. Patil, M. A. S. R. Saadi, T. Xie, J. G. Azadani, B. Uluutku, R. Amin, A. F. Briggs, M. Miscuglio, D. Van Thourhout, S. D. Solares, T. Low, R. Agarwal, S. R. Bank, and V. J. Sorger, “Strain Engineered high responsivity MoTe2 photodetectors for Si photonic integrated circuits,” Nat. Photonics (2020). [CrossRef]  

20. P. Ma, N. Flöry, Y. Salamin, B. Baeuerle, A. Emboras, A. Josten, T. Taniguchi, K. Watanabe, L. Novotny, and J. Leuthold, “Fast MoTe2 Waveguide Photodetector with High Sensitivity at Telecommunication Wavelengths,” ACS Photonics 5(5), 1846–1852 (2018). [CrossRef]  

21. N. Flöry, P. Ma, Y. Salamin, A. Emboras, T. Taniguchi, K. Watanabe, J. Leuthold, and L. Novotny, “Waveguide-Integrated Van der Waals Heterostructure Photodetector at Telecom Wavelengths with High Speed and High Responsivity,” Nat. Nanotechnol. 15(2), 118–124 (2020). [CrossRef]  

22. Y. Ding, Z. Cheng, X. Zhu, K. Yvind, J. Dong, M. Galili, H. Hu, N. A. Mortensen, S. Xiao, and L. K. Oxenløwe, “Ultra-compact integrated graphene plasmonic photodetector with bandwidth above 110 GHz,” Nanophotonics 9(2), 317–325 (2020). [CrossRef]  

23. M. Long, E. Liu, P. Wang, A. Gao, H. Xia, W. Luo, B. Wang, J. Zeng, Y. Fu, K. Xu, W. Zhou, Y. Lv, S. Yao, M. Lu, Y. Chen, Z. Ni, Y. You, X. Zhang, S. Qin, Y. Shi, W. Hu, D. Xing, and F. Miao, “Broadband Photovoltaic Detectors Based on an Atomically Thin Heterostructure,” Nano Lett. 16(4), 2254–2259 (2016). [CrossRef]  

24. G. W. Mudd, S. A. Svatek, L. Hague, O. Makarovsky, Z. R. Kudrynskyi, C. J. Mellor, P. H. Beton, L. Eaves, K. S. Novoselov, Z. D. Kovalyuk, E. E. Vdovin, A. J. Marsden, N. R. Wilson, and A. Patanè, “High Broad-Band Photoresponsivity of Mechanically Formed InSe–Graphene van der Waals Heterostructures,” Adv. Mater. 27(25), 3760–3766 (2015). [CrossRef]  

25. W. Feng, J. Wu, X. Li, W. Zheng, X. Zhou, K. Xiao, W. Cao, B. Yang, J. Idrobo, L. Basile, W. Tian, P. Tan, and P. Hu, “Ultrahigh photo-responsivity and detectivity in multilayer InSe nanosheets phototransistors with broadband response,” J. Mater. Chem. C 3(27), 7022–7028 (2015). [CrossRef]  

26. E. Liu, M. Long, J. Zeng, W. Luo, Y. Wang, Y. Pan, W. Zhou, B. Wang, W. Hu, Z. Ni, Y. You, X. Zhang, S. Qin, Y. Shi, K. Watanabe, T. Taniguchi, H. Yuan, H. Y. Hwang, Y. Cui, F. Miao, and D. Xing, “High Responsivity Phototransistors Based on Few-Layer ReS2 for Weak Signal Detection,” Adv. Funct. Mater. 26(12), 1938–1944 (2016). [CrossRef]  

27. D. A. Bandurin, A. V. Tyurnina, G. L. Yu, A. Mishchenko, V. Zólyomi, S. V. Morozov, R. K. Kumar, R. V. Gorbachev, Z. R. Kudrynskyi, S. Pezzini, Z. D. Kovalyuk, U. Zeitler, K. S. Novoselov, A. Patanè, L. Eaves, I. V. Grigorieva, V. I. Fal’ko, A. K. Geim, and Y. Cao, “High electron mobility, quantum Hall effect and anomalous optical response in atomically thin InSe,” Nat. Nanotechnol. 12(3), 223–227 (2017). [CrossRef]  

28. X. Ling, H. Wang, S. Huang, F. Xia, and M. S. Dresselhaus, “The renaissance of black phosphorus,” Proc. Natl. Acad. Sci. 112(15), 4523–4530 (2015). [CrossRef]  

29. S. Kim, A. Konar, W. S. Hwang, J. H. Lee, J. Lee, J. Yang, C. Jung, H. Kim, J. B. Yoo, J. Y. Choi, Y. W. Jin, S. Y. Lee, D. Jena, W. Choi, and K. Kim, “High-Mobility and Low-Power Thin-Film Transistors Based on Multilayer MoS2 Crystals,” Nat. Commun. 3(1), 1011 (2012). [CrossRef]  

30. S. -L. Li, K. Komatsu, S. Nakaharai, Y.-F. Lin, M. Yamamoto, X. Duan, and K. Tsukagoshi, “Thickness scaling effect on interfacial barrier and electrical contact to two-dimensional MoS2 layers,” ACS Nano 8(12), 12836–12842 (2014). [CrossRef]  

31. L. Yang, K. Majumdar, H. Liu, Y. Du, H. Wu, M. Hatzistergos, P. Y. Hung, R. Tieckelmann, W. Tsai, C. Hobbs, and P. D. Ye, “Chloride molecular doping technique on 2D materials: WS2 and MoS2,” Nano Lett. 14(11), 6275–6280 (2014). [CrossRef]  

32. R. Kappera, D. Voiry, S. E. Yalcin, B. Branch, G. Gupta, A. D Mohite, and M. Chhowalla, “Phase-engineered low-resistance contacts for ultrathin MoS2 transistors,” Nat. Mater. 13(12), 1128–1134 (2014). [CrossRef]  

33. D. J. Frank, Y. Taur, and H. -. P. Wong, “Generalized scale length for two-dimensional effects in MOSFETs,” IEEE Electron Device Lett. 19(10), 385–387 (1998). [CrossRef]  

34. S. A. Maier, Plasmonics: Fundamentals and Applications (Springer, 2007).

35. J. D. Joannopoulos, R. D. Meade, and J. N. Winn, Photonic Crystals (Princeton Univ. Press, 1995).

36. Z. Ma, M. Tahersima, S. Khan, and V. J. Sorger, “2D material-based mode overlap engineering in electro-optic modulators,” IEEE J. Sel. Top. Quantum Electron. 23(1), 81–88 (2017). [CrossRef]  

37. H. Yoshida, Y. Yamashita, M. Kuwabara, and H. Kan, “A 342-nm ultraviolet AlGaN multiple-quantum-well laser diode,” Nat. Photonics 2(9), 551–554 (2008). [CrossRef]  

38. S. Takagi, S.-H. Kim, M. Yokoyama, K. Nishi, R. Zhang, and M. Takenaka, “Material Challenges and Opportunities in Ge/III-V channel MOSFETs,” ECS Trans. 64(11), 99–110 (2014). [CrossRef]  

39. M. Houssa, E. Chagarov, and A. Kummel, “Surface defects and passivation of Ge and III-V interfaces,” MRS Bull. 34(7), 504–513 (2009). [CrossRef]  

40. J. E. Muench, A. Ruocco, M. A. Giambra, V. Miseikis, D. Zhang, J. Wang, H. F. Y. Watson, G. C. Park, S. Akhavan, V. Sorianello, M. Midrio, A. Tomadin, C. Coletti, M. Romagnoli, A. C. Ferrari, and I. Goykhman, “Waveguide-Integrated, Plasmonic Enhanced Graphene Photodetectors,” Nano Lett. 19(11), 7632–7644 (2019). [CrossRef]  

41. Z. Ma, K. Kikunaga, H. Wang, S. Sun, R. Amin, R. Maiti, M. H. Tahersima, H. Dalir, M. Miscuglio, and V. J. Sorger, “Compact Graphene Plasmonic Slot Photodetector on Silicon-on-insulator with High Responsivity,” ACS Photonics 7(4), 932–940 (2020). [CrossRef]  

42. J. Michel, J. Liu, and L. C. Kimerling, “High-performance Ge-on-Si photodetectors,” Nat. Photonics 4(8), 527–534 (2010). [CrossRef]  

43. T. Mueller and E. Malic, “Exciton physics and device application of two-dimensional transition metal dichalcogenide semiconductors,” npj 2D Mater. Appl. 2(1), 29 (2018). [CrossRef]  

44. R. Maiti, R. A. Hemnani, R. Amin, Z. Ma, M. Tahersima, T. A. Empante, H. Dalir, R. Agarwal, L. Bartels, and V. J. Sorger, “A semi-empirical integrated microring cavity approach for 2D material optical index identification at 1.55 um,” Nanophotonics 8(3), 435–441 (2019). [CrossRef]  

45. X. Cui, G.-H. Lee, Y. D. Kim, G. Arefe, P. Y. Huang, C.-H. Lee, D. A. Chenet, X. Zhang, L. Wang, F. Ye, F. Pizzocchero, B. S. Jessen, K. Watanabe, T. Taniguchi, D. A. Muller, T. Low, P. Kim, and J. Hone, “Multi-terminal transport measurements of MoS2 using a van der Waals heterostructure device platform,” Nat. Nanotechnol. 10(6), 534–540 (2015). [CrossRef]  

46. H. Shi, R. Yan, S. Bertolazzi, J. Brivio, B. Gao, A. Kis, D. Jena, H. G. Xing, and L. Huang, “Exciton dynamics in suspended mono layer and few-layer MoS2 2D crystals,” ACS Nano 7(2), 1072–1080 (2013). [CrossRef]  

47. https://www.imec-int.com/en/imec-magazine/imec-magazine-february-2019/a-300mm-platform-for-2d-material-based-mosfet-devices

48. T.-A. Chen, C. -P. Chuu, C. -C. Tseng, C. -K. Wen, H. -S. P. Wong, S. Pan, R. Li, T. -A. Chao, W.-C. Chueh, Y. Zhang, Q. Fu, B. I. Yakobson, W. -H. Chang, and L. -J. Li, “Wafer-scale single-crystal hexagonal boron nitride monolayers on Cu (111),” Nature 579(7798), 219–223 (2020). [CrossRef]  

49. A. C. Gomez, M. Buscema, R. Molenaar, V. Singh, L. Janssen, H. S. J. van der Zant, and G. A. Steele, “Deterministic transfer of two-dimensional materials by all-dry viscoelastic stamping,” 2D Mater. 1(1), 011002 (2014). [CrossRef]  

50. R. Maiti, C. Patil, R. Hemnani, M. Miscuglio, R. Amin, Z. Ma, R. Chaudhary, A. T. C. Johnson, L. Bartels, R. Agarwal, and V. J. Sorger, “Loss and Coupling Tuning via Heterogeneous Integration of MoS2 Layers in Silicon Photonics,” Opt. Mater. Express 9(2), 751–759 (2019). [CrossRef]  

51. R. A. Hemnani, C. Carfano, J. P. Tischler, M. H. Tahersima, R. Maiti, L. Bartels, R. Agarwal, and V. J. Sorger, “Towards a 2D Printer: A Deterministic Cross Contamination-free Transfer Method for Atomically Layered Materials,” 2D Mater. 6(1), 015006 (2019). [CrossRef]  

52. K. He, C. Poole, and K. F. Mak, “Experimental demonstration of continuous electronic structure tuning via strain in atomically thin MoS2,” Nano Lett. 13(6), 2931–2936 (2013). [CrossRef]  

53. H. J. Conley, B. Wang, and J. I. Ziegler, “Bandgap engineering of strained monolayer and bilayer MoS2,” Nano Lett. 13(8), 3626–3630 (2013). [CrossRef]  

54. C R Zhu, G Wang, and B L. Liu, “Strain tuning of optical emission energy and polarization in monolayer and bilayer MoS2,” Phys. Rev. B 88(12), 121301 (2013). [CrossRef]  

55. A. Lipatov, H. Lu, M. Alhabeb, B. Anasori, A. Gruverman, Yury Gogotsi, and A. Sinitskii, “Elastic properties of 2D Ti3C2Tx MXene monolayers and bilayers,” Sci. Adv. 4(6), eaat0491 (2018). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (1)

Fig. 1.
Fig. 1. Design strategies for Engineering Gain-Bandwidth-Product (GBP) Photodetectors. a) Photodetector performance cluster into two quadrants; high responsivity yet low speed (bandwidth, BW), and vice versa. Target detectors with a GBP of ∼1012 - 1013 scale orthogonally (upper-right quadrant) to the iso-GBP lines. The values are taken from [526]. b) Plot of bandwidth vs. photocarrier-collecting electrical channel length showing two regimes where the transit time-limited bandwidth (BWtransit) is dominant for long-channel detectors (gray shade), whereas for sub 100 nm channel lengths, the bandwidth is limited by RC time (blue shade). BWtransit is estimated for four different mobility values (1 - 1000 cm2/Vs) for a Vsd = 1 V taken from Ref. [19, 2729]. The RC time-limited bandwidth (BWRC) is estimated for four different resistance values starting from 0.1−100 kΩ [3032], where, the electrical capacitance is determined by a parallel-plate model (fringe fields ignored) for a lateral junction. A finding of this parametric study is that for a (arbitrarily selected) target speed of 100 GHz; i) micrometer long channel-based photodetectors required a rather high mobility (<104 cm2/Vs) but with a somewhat relaxed contact resistance RC requirements, yet high mobilities challenges achievable resistance due to scattering events induced by higher doping, while ii) 10’s nm short-channel detectors are only resistance limited, since (near) ballistic transport is given even for poor mobilities (<10 cm2/Vs), and relatively high grain can be achieved by reducing the transit time to ∼ps as supposed to increasing the carrier lifetime. Such short-channel performance-detectors, however, demand optical mode squeezing to adhere to device scale-length theory (SLT) rules which can be achieved with slot-waveguide designs.
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.