Towards practical carrier multiplication: Donor ... - ACS Publications

Kistagången 16, 164 40 Kista, Sweden. 2National Renewable Energy Laboratory,. 15013 Denver West Parkway, Golden, CO 80401, 303-275-3000, USA...
0 downloads 0 Views 1MB Size
Article Cite This: ACS Photonics 2018, 5, 2843−2849

Toward Practical Carrier Multiplication: Donor/Acceptor Codoped Si Nanocrystals in SiO2 Nguyen Xuan Chung,†,‡ Rens Limpens,*,‡,§ Chris de Weerd,∥ Arnon Lesage,∥ Minoru Fujii,⊥ and Tom Gregorkiewicz*,∥ †

Royal Institute of Technology (KTH) Kistagången 16, 164 40 Kista, Sweden National Renewable Energy Laboratory, 15013 Denver West Parkway, Golden, Colorado 80401, United States ∥ Van der Waals-Zeeman Institute, University of Amsterdam, Science Park 904, 1098 XH Amsterdam, The Netherlands ⊥ Department of Electrical and Electronic Engineering, Kobe University, Rokkodai, Nada, Kobe 657-8501, Japan

Downloaded via KAOHSIUNG MEDICAL UNIV on November 16, 2018 at 06:04:46 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

§

S Supporting Information *

ABSTRACT: Carrier multiplication (CM) is an interesting fundamental phenomenon with application potential in optoelectronics and photovoltaics, and it has been shown to be promoted by quantum confinement effects in nanostructures. However, mostly due to the short lifetimes of additional electron−hole (e-h) pairs generated by CM, major improvements of quantum dot devices that exploit CM are limited. Here we investigate CM in SiO2 solid state dispersions of phosphorus and boron codoped Si nanocrystals (NCs): an exotic variant of Si NCs whose photoluminescence (PL) emission energy, the optical bandgap, is significantly red-shifted in comparison to undoped Si NCs. By combining the results obtained by ultrafast induced absorption (IA) with PL quantum yield (PL QY) measurements, we demonstrate CM with a long (around 100 μs) lifetime of the additional e-h pairs created by the process, similar as previously reported for undoped Si NCs, but with a significantly lower CM threshold energy. This constitutes a significant step toward the practical implementation of Si-based NCs in optoelectronic devices: we demonstrate efficient CM at the energy bandgap optimal for photovoltaic conversion. KEYWORDS: carrier multiplication, carrier generation rate, codoped silicon nanocrystals, photoluminescence quantum yield

T

bottleneck remains the utilization of the additional e-h pairs. Their extraction (as well as radiative recombination) is dramatically hindered by the efficient Auger recombination (AR, ps-range), which is also enhanced in nanostructures by quantum confinement. While it has been shown that the additional e-h pairs generated by CM can escape rapid AR by dedicated device optimization of PbS NCs14 and hydrazine surface treated PbSe NC films,11 the problem is severe. In this study we address both challenges by optimizing the CM threshold and the extraction feasibility at the same time. We do this by making use of P and B codoped Si NCs. Previously we have demonstrated that AR can be suppressed in dense solid state dispersions of Si NCs in SiO2 by spatially separating the eh pairs into neighboring NCs,15 in analogy to the singlet fission process in organic chromophores.16 This is the essential reason why we picked Si NCs for this study. At the same time, we expect that, by P and B codoping, the threshold energy of CM might decrease, as a consequence of the optical bandgap reduction, while still maintaining the single-exciton recombi-

he quest for realization of efficient carrier multiplication (CM) has been the primary goal of many investigations in nanomaterials and, specifically, in semiconductor nanocrystals (NCs) for more than a decade.1−5 From the point of view of potential application in photovoltaics, the most important aspects are the (i) energy of the electron−hole (e-h) pairs and the (ii) threshold energy for CM. Indeed, for the solar spectrum AM 1.5, the optimal e-h pair energy for CM yielding the highest conversion efficiency of approximately 40% is around 0.8 eV.6 This low energy cannot be attained in Si NCs, and that partly explains the wide investigation of CM in toxic Pb compounds, such as PbS and PbSe; we do think it is important to mention the low bandgap energy allotropes of Si.7,8 Experimental investigations of CM in these materials have not been reported, to our knowledge, but would certainly be interesting. To this end, we note that while some investigations focus on band structure engineering in order to promote the CM process itself by, for example, making use of core−shell NCs,9,10 others try to optimize the harvesting of the additional e-h pairs generated by CM.11 The latter is essential since, whereas increased Coulomb interactions and decelerated hotcarrier cooling rates significantly enhance the CM potential, as has been shown for multiple NC systems,1,2,4,12,13 the main © 2018 American Chemical Society

Received: February 2, 2018 Published: May 17, 2018 2843

DOI: 10.1021/acsphotonics.8b00144 ACS Photonics 2018, 5, 2843−2849

Article

ACS Photonics nation of the additional e-h pairs that are created by the process. Doping of NCs has been investigated as an interesting option to influence their optical and electronic properties.17 While mostly explored for one type of dopants only,18,19 the simultaneous doping with electron and hole donating species, hereby creating neutrally charged nanoparticles, is particularly interesting. This has been explored for Si NCs and shown to create new possibilities in terms of the surface functionalization,20 dispersibility,21,22 and electronic capabilities.23−25 One specific advantage was the lowering of the optical bandgap,26,27 below that of bulk Si, facilitated by the radiative recombination between the donor and acceptor states that are introduced by the P/B doping, as recently confirmed by a scanning tunneling microscopy study.28 As explained before, this makes codoped Si NCs interesting for CM, and in our preliminary report, we have indeed presented a fingerprint of CM in these materials.29 In light of this study it is important to realize that CM efficiencies are traditionally inferred from ultrafast transient absorption techniques, which monitor the fast nonradiative recombination of the multiple excitons by Auger recombination.30 As shown previously,31 this method fails in case of closely packed SiO2 embedded Si NCs due to the delocalized nature of the CM process. Instead, here we derive the carrier generation efficiency by rigorously scaling the number of photons emitted in photoluminescence (PL)15 and the induced absorption (IA) amplitude31 to the number of absorbed photons, analogously to THz investigations of CM in bulk PbS and PbSe.4 In this study we cross correlate both the IA and the PL approach to quantify the carrier generation rate and to investigate the carrier dynamics of the additional e-h pairs in the codoped Si NCs. In that way, the CM efficiency at particular excitation energies is found and the nature of CM and its threshold energy can be determined.

in amplifier. Furthermore, linear absorption has been measured with a dual beam mode PerkinElmer Lambda 950 spectrometer. IA measurements are performed at two different systems. To reach the nonlinear excitation regime and verify the singleexciton dynamics the pump−probe data is collected using a femtosecond transient absorption spectrometer (Helios, Ultrafast Systems). The laser source is a 4 W Ti-sapphire amplifier (Libra, Coherent), operating at 1 kHz and 100 fs pulse width (see ref 33 for further details). To investigate the carrier generation rates, a 1028 nm fundamental laser is produced by Yb:KGW oscillator (Light Conversion, Pharos SP) with a repetition rate of 5 kHz and a duration of 180 fs and then is split into two beams. The first beam is nonlinear mixed in an Optical Parametric Amplifier (OPA) and passes a second harmonic module (Light Conversion, Orpheus), pump beam with varied wavelengths between 310 and 1500 nm is produced. Being modulated by a chopper, the repetition frequency of the pump beam is 2.5 kHz. The second beam passes a sapphire crystal and produces a broadband probe (500−1600 nm). The probe spectra are recorded on a CCD (Ultrafast Systems, Helios). Relative PL quantum yield (QY) of the codoped sample is measured by using two different setups in combination with two excitation sources. The sample is excited inside a Newport integrating sphere (with a diameter of 7.5 cm) in order to collect emitted photons in all directions and to avoid the scattering effects. For excitation energies above 2.1 eV, a highly stable L2272 Hamamatsu xenon lamp has been used as excitation source. For excitation energies below 2.3 eV, because of low absorption and low emission in this region, an optical parameter oscillator (OPO) producing a varied wavelength laser with high power is employed as an excitation source. In order to evaluate the number of emitted photons, the same combination of the Edinburgh germanium detector EI-A with the DSP 7280 lock-in amplifier is used.

MATERIALS AND METHODS Materials. The thin films containing Si NC samples are deposited by a magnetron RF cosputtering system, with a consequent annealing process as follows. Full details of the production procedure can be found in ref.32 Two samples are chosen for this study. P and B codoped NCs prepared by sputtering a phosphosilicate glass target partially covered with pieces of silicon and B2O3 tablets. An undoped reference Si NC sample is produced by using a SiO2 target covered with pieces of silicon. For each sample, a layer of approximately 1.5 μm is deposited on top of a fused silica substrate. The samples are then annealed for 30 min at 1150 °C in nitrogen atmosphere. This high-temperature heat treatment induces phase separation between Si and SiO2, and Si NCs (either un- or codoped) are formed. Based on this same previous work, also the concentration of dopants could be determined for the used preparation procedure and is determined to contain 1.25% atomic concentration of both P and B. The excess silicon concentration is set in both samples at 18%. Methods. Standard PL spectra are investigated by making use of the 3.64 eV excitation of a xenon lamp coupled to a Solar LS MSA130 monochromator. For the pure Si NCs, the PL spectrum is detected with a Hamamatsu CCD (S7031−1008S) integrated in a Solar LS M266 monochromator, whereas the near-infrared (NIR) PL spectrum of the codoped NCs is investigated on a setup including a Spex 1000 M monochromator with a high resolution grating for the NIR detection, an Edinburgh EI-A germanium detector and a DSP 7280 lock-

RESULTS AND DISCUSSION Sample Characterization. Before investigating CM, we start with the optical characterization of the samples. In Figure 1 we present PL and absorption spectra of the P and B codoped Si NCs and compare them to those of the undoped NCs as a reference. In agreement with previous reports,34,35 we confirm that the codoping significantly red-shifts the optical bandgap to a value below the bandgap of bulk silicon (i.e., 1.12 eV). At the same time, the PL lifetime is of the order of 100 μs.20 The spectral downshift of 0.45 eV is ascribed to donor-to-acceptor transitions in fully compensated NCs,34 and is perfectly in line with a recent scanning tunneling microscopy (STM) study of similar sized codoped Si NCs ref.28 Here it is shown that these codoped Si NCs with donor−acceptor transitions of around 1 eV have an intrinsic bandgap of around 1.5 eV (similar to the undoped NCs in this work). We further note considerable (≈2.5%) absorption in the codoped sample for below-band gap energies. This we ascribe to “parasitic” absorption (i.e., not contributing to PL), which most probably originates from the excitation of free carriers in not fully compensated NCs (as discussed further on), with a possible contribution of defect states unavoidably introduced by doping. Furthermore, the overall linear absorption in the VIS regime is significantly enhanced in the codoped NCs, which is related to a higher Si excess and, consequently, a somewhat higher concentration of Si NCs in the codoped sample. Ultrafast Carrier Dynamics. In Figure 2a we display IA decay transients for the codoped NCs, for pump energies from





2844

DOI: 10.1021/acsphotonics.8b00144 ACS Photonics 2018, 5, 2843−2849

Article

ACS Photonics

Consequently, the absence of an ultrafast decaying component for high energy pumping indicates an absence of AR, and therefore also the absence of the CM process in the traditional sense of the term. We therefore conclude that we do not observe carrier multiplication within the same NC (also referred to as multiple exciton generation). In the next part, by monitoring the carrier generation yield, we investigate the potential occurrence of the spatially separated form of CM. In order to analyze the IA dynamics further we recall that oxidized Si NCs typically suffer from efficient trapping of free carriers, with the effect being enhanced at higher energies, that is, for “hot” carriers,30,36,37 thus, quenching the IA amplitude on the time-scales of tens of picoseconds. As a result, even at pump energies below 2 Eg (excluding possible CM) and in linear excitation regime with Nabs ≪ 1 (so under conditions precluding a possibility of generation of multiple excitons in a single NC) the carriers are prone to nonradiative recombination. This manifests itself by the nonunity baseline of the A-toB ratios for low pump power densities, depicted in the inset of Figure 2b (for pump energy of 2.27 eV). In our analysis, we take the intensity of the IA traces at 1 ps (defined as A) and compare it to the intensity at 500 ps (B). Furthermore, as the number of absorbed photons is increased (by changing the pump power) we observe increasing A-to-B ratios - being the result of Auger-induced quenching of the multiple excitons that are created by multiple-photon-absorptions in the same NC. This behavior independently verifies that all traces in Figure 2a are representing the single-exciton regime. Photoluminescence Quantum Yield. Relative PL QY of the codoped Si NCs has been investigated in two different setups whose details are given in the Supporting Information. For the measurements with excitation energies below 2.3 eV (depicted by the orange symbols), the experimental error is governed by the laser power fluctuations of 5−10%. Moreover, the PL QY values measured in that range have to be corrected for the strong parasitic absorption, see Figure 1. The correction procedure is discussed in detail in the Supporting Information. As such, and in combination with the rather low absorption cross section at excitation energies below Eexc = 2 eV (as a

Figure 1. Normalized PL spectra of undoped (thick cyan) and P and B codoped (thick red curve) Si NCs upon an excitation energy of 3.64 eV. Normalization is performed by scaling the PL maxima to one. The dark red curve represents a log-normal fit of the PL spectrum of the codoped NCs. Absorption spectra of the undoped and the codoped NCs (thin cyan and thin red curves, respectively) are investigated in the excitation energy region below 2 eV. Parasitic absorption (evaluated by the level of absorption in the NIR regime) of the pure (cyan dashed line) and codoped (red dashed line) NCs is found to be around 0.3 and 2.5%, respectively. The missing regions of absorption spectra indicate a detector switch.

below to above two times the optical bandgap (Eg). These have been obtained in the linear regime, so with significantly less than one photon absorbed per NC, Nabs ≪ 1. The traces are averaged over a range of probing wavelengths (1200−1350 nm) to increase the signal-to-noise ratio - this is possible since no probing energy dependence is detected in this domain. The traces are normalized at 3 ps. As can be concluded, identical IA dynamics are obtained regardless of the excitation energy, below and well above 2 Eg, that is, where eventual CM might be expected. As mentioned in the first section, the existence of CM is traditionally inferred from a change in IA dynamics, induced by rapid AR of multiple excitons colocalized in the same NC.

Figure 2. (a) Normalized IA traces of the codoped Si NCs, for pump energies below and above 2 Eg, (the lowest possible threshold energy for CM, at Nabs ≪ 1). The traces are averaged over a range (1200−1350 nm) of probing wavelengths. (b) IA traces for three different excitation power densities, ranging from 21 to 127 mW/cm2, for a pump energy of 2.27 eV. The inset shows the derived A-to-B ratios (representing the ratio of the trace amplitude at 1 and 500 ps, respectively) as a function of the excitation power density, averaged over a probing range of 1100−1200 nm, verifying the linear excitation regime (Nabs ≪ 1). Note that the nonunity baseline of the A/B ratio is indicative of ultrafast nonradiative recombination channels, in the absence of AR. 2845

DOI: 10.1021/acsphotonics.8b00144 ACS Photonics 2018, 5, 2843−2849

Article

ACS Photonics

value around 5%, as shown in the Supporting Information. This low PL QY we attribute to the unavoidably large fraction of not perfectly compensated NCs where the nonradiative Auger recombination will prevail. While compensated doping, with an equal number of donors and acceptors per NC is energetically favorable,41 this advantage of formation energy diminishes for a higher number of dopants per NC. Therefore, within our sample (where the average nominal number of P and B dopants per NC is relatively high,42 >10), we do expect a significant amount of uncompensated NCs. Although it is at the moment unclear what the actual fraction of uncompensated NCs should be (especially so, since the number of electrically active dopants is lower than the nominal doping level), the large drop of the PL QY (∼factor 5 with our reference sample) suggest a significant fraction. As a result, while the radiative recombination will dominate fully for compensated NCs, it will be quenched in those with unbalanced doping. Carrier Generation Rate. In Figure 4, we combine the carrier generation rates calculated from both the PL QY and the IA measurements. The rate is defined as the number of e-h pairs created upon the absorption of a single photon. The baseline values of the absolute PL QY for pump energies below 2 Eg (which are presented in Figure 3) are normalized to 1. To derive the carrier generation rate from the IA dynamics we normalized all traces to the number of absorbed photons at that specific pump condition, in a similar way to our previous study for pure Si NCs,31 and used the amplitude (at 3 ps) from the normalized traces as the relative carrier generation rate. To optimize the result, we used three different excitation powers (all in the linear regime), for each pump energy, following ref 43; the details of this procedure can be found in the Supporting Information. These relative IA carrier generation rates scaled to those derived from the PL QY are presented in Figure 4. Detailed Interpretation. The first observation we make from Figure 4 is that the carrier generation rates as determined from IA and PL QY data are identical. Since both techniques measure the free carrier population at very different time scales of ps and near ms, respectively, this implies that the free carrier population at long delay times is proportional to that shortly after the pump pulse, for all excitation energies. We can therefore exclude that the rise of the PL QY at high excitation energies is resulting from decreased nonradiative recombination channels; if that would be the case, the change in the

result of the indirect bandgap of Si), it is extremely challenging to perform PL QY measurements at low excitation energies. We direct the interested reader to a recent paper of ours (ref 38). In spite of these limitations and forthcoming larger scatter of carrier yield data we managed to go as low as Eexc = 1.5 eV (Figure 3). For the high excitation energies above 2.1 eV (blue

Figure 3. Absolute PL QY of the codoped Si NCs, measured by using a xenon lamp (blue circles) and a laser (orange triangles) as excitation sources, as described in the Methods. The error bars for the data obtained under laser excitation are larger due to the significant power fluctuations and the low linear absorption in this energy range.

symbols), the linear absorption of the NCs is fairly high, and a xenon lamp can be used; its power fluctuations are low (around 0.25%), reducing the error bars to within the symbol size. In that range also the absolute value of the PL QY can be determined; see ref 29 for details. We end up having two sets of partially overlapping PL QY data, with the one at high excitation energy containing information on the absolute PL QY. Therefore, by rescaling, the relative PL QY in the low excitation energy range can be converted to the absolute values. The results are shown in Figure 3. Inspecting Figure 3, we note that (i) the PL QY increases for higher excitation energies and (ii) the absolute PL QY is quite low, between 0.5 and 1.5%, indicating that the optical activity of the ensemble of codoped Si NCs is considerably lower than 5−35% typically observed for undoped NCs,15,38−40 with the reference sample featuring a

Figure 4. (a) Carrier generation yield in the codoped Si NCs, obtained from PL QY (black markers) and IA (red markers) measurements. The effect of ideal CM (i.e., commencing at the energy conservation limit and with 100% efficiency,15 see text for further explanation) is presented by the blue curve. The error bars of the IA-based data are derived from the noise level of the IA amplitude at 3 ps and the uncertainty in the estimation of the number of absorbed photons per NC. (b) Magnification of panel (a) for excitation energies below 2.5 eV. 2846

DOI: 10.1021/acsphotonics.8b00144 ACS Photonics 2018, 5, 2843−2849

Article

ACS Photonics

the respective contributions of these specific fractions of the total NC ensemble, they obviously affect the optical activity, being responsible for the relatively low PL QY yield observed in our materials. However, it is important to realize that this does not influence the experimental method used to determine the delocalized form of carrier multiplication, since the effect of these different fractions on the ensemble optical properties is already accounted for in the “baseline” carrier generation efficiency, for excitation energies below 2 times the optical bandgap ( 2.5 eV, the experimentally measured carrier generation is considerably lower than predicted by the simulation. Obviously in that range detrimental processes leading to carrier loss (such as, e.g., the increased role of highenergy traps) appear, effectively counterbalancing the positive effect of CM, as discussed in ref 38. (ii) As mentioned, the investigated doped NC ensemble consists of a combination of compensated, uncompensated, and doping-induced defected NCs. Although based on the results at hand we cannot separate



CONCLUSIONS



ASSOCIATED CONTENT

In conclusion, by combining results obtained by calibrated ultrafast induced absorption and photoluminescence spectroscopies we demonstrated a delocalized form of CM in closepacked P and B codoped Si NCs in SiO2. Similar as for SiO2 dispersions of undoped Si NCs, the additional e-h pairs occupy adjacent NCs and hereby escape rapid nonradiative Auger recombination. This is confirmed by the absence of an Auger component in the ultrafast carrier dynamics, while the carrier generation efficiency shows a step-like increase at multiples of the optical bandgap. As such, multiple excitons are not subject to AR, allowing for long-extraction times (>100 μs). This is beneficial for optoelectronic applications and is in contrast to the ps extraction-window typical for CM-generated carriers in traditional CM. Moreover, the donor and acceptor levels in these codoped nanoparticles significantly decrease the CM threshold to 2 Eg, to below hν ≈ 2 eV on the absolute energy scale. Therefore, the current results show that more than half of all photons of the solar spectrum are available for CM.

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsphotonics.8b00144. The setup to measure the PL QY, transmission, relative PL QY, absolute PL QY, and the parasitic absorption. In addition, we provide IA dynamics for six excitation energies (PDF).



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. *E-mail: [email protected]. ORCID

Rens Limpens: 0000-0002-2417-9389 Chris de Weerd: 0000-0002-8826-2616 Minoru Fujii: 0000-0003-4869-7399 Tom Gregorkiewicz: 0000-0003-2092-8378 Author Contributions ‡

These authors contributed equally to this work.

2847

DOI: 10.1021/acsphotonics.8b00144 ACS Photonics 2018, 5, 2843−2849

Article

ACS Photonics Author Contributions

(15) Timmerman, D.; Valenta, J.; Dohnalova, K.; de Boer, W. D.; Gregorkiewicz, T. Step-like enhancement of luminescence quantum yield of silicon nanocrystals. Nat. Nanotechnol. 2011, 6 (11), 710−713. (16) Smith, M. B.; Michl, J. Singlet fission. Chem. Rev. 2010, 110 (11), 6891−6936. (17) Norris, D. J.; Efros, A. L.; Erwin, S. C. Doped nanocrystals. Science 2008, 319 (5871), 1776−1779. (18) Luther, J. M.; Jain, P. K.; Ewers, T.; Alivisatos, A. P. Localized surface plasmon resonances arising from free carriers in doped quantum dots. Nat. Mater. 2011, 10 (5), 361−366. (19) Mocatta, D.; Cohen, G.; Schattner, J.; Millo, O.; Rabani, E.; Banin, U. Heavily doped semiconductor nanocrystal quantum dots. Science 2011, 332 (6025), 77−81. (20) Sugimoto, H.; Fujii, M.; Imakita, K.; Hayashi, S.; Akamatsu, K. Codoping n-and p-type impurities in colloidal silicon nanocrystals: Controlling luminescence energy from below bulk band gap to visible range. J. Phys. Chem. C 2013, 117 (22), 11850−11857. (21) Fukuda, M.; Fujii, M.; Sugimoto, H.; Imakita, K.; Hayashi, S. Surfactant-free solution-dispersible Si nanocrystals surface modification by impurity control. Opt. Lett. 2011, 36 (20), 4026−4028. (22) Sugimoto, H.; Fujii, M.; Fukuda, Y.; Imakita, K.; Akamatsu, K. All-inorganic water-dispersible silicon quantum dots: highly efficient near-infrared luminescence in a wide pH range. Nanoscale 2014, 6 (1), 122−126. (23) Gutsch, S.; Laube, J.; Hiller, D.; Bock, W.; Wahl, M.; Kopnarski, M.; et al. Electronic properties of phosphorus doped silicon nanocrystals embedded in SiO2. Appl. Phys. Lett. 2015, 106 (11), 113103. (24) Nakamura, T.; Adachi, S.; Fujii, M.; Sugimoto, H.; Miura, K.; Yamamoto, S. Size and dopant-concentration dependence of photoluminescence properties of ion-implanted phosphorus-and boroncodoped Si nanocrystals. Phys. Rev. B: Condens. Matter Mater. Phys. 2015, 91 (16), 165424. (25) Higashikawa, Y.; Azuma, Y.; Majima, Y.; Kano, S.; Fujii, M. Integration of colloidal silicon nanocrystals on metal electrodes in single-electron transistor. Appl. Phys. Lett. 2016, 109 (21), 213104. (26) Fujii, M.; Yamaguchi, Y.; Takase, Y.; Ninomiya, K.; Hayashi, S. Photoluminescence from impurity codoped and compensated Si nanocrystals. Appl. Phys. Lett. 2005, 87 (21), 211919. (27) Hori, Y.; Kano, S.; Sugimoto, H.; Imakita, K.; Fujii, M. SizeDependence of Acceptor and Donor Levels of Boron and Phosphorus Codoped Colloidal Silicon Nanocrystals. Nano Lett. 2016, 16 (4), 2615−2620. (28) Ashkenazi, O.; Azulay, D.; Balberg, I.; Kano, S.; Sugimoto, H.; Fujii, M.; et al. Size-dependent donor and acceptor states in codoped Si nanocrystals studied by scanning tunneling spectroscopy. Nanoscale 2017, 9 (45), 17884−17892. (29) Chung, N. X.; Limpens, R.; Lesage, A.; Fujii, M.; Gregorkiewicz, T. Optical generation of electron−hole pairs in phosphor and boron co-doped Si nanocrystals in SiO2. Phys. Status Solidi A 2016, 213 (11), 2863−2866. (30) Trinh, M. T.; Limpens, R.; Gregorkiewicz, T. Experimental investigations and modeling of Auger recombination in silicon nanocrystals. J. Phys. Chem. C 2013, 117 (11), 5963−5968. (31) Trinh, M. T.; Limpens, R.; De Boer, W. D.; Schins, J. M.; Siebbeles, L. D.; Gregorkiewicz, T. Direct generation of multiple excitons in adjacent silicon nanocrystals revealed by induced absorption. Nat. Photonics 2012, 6 (5), 316−321. (32) Fujio, K.; Fujii, M.; Sumida, K.; Hayashi, S.; Fujisawa, M.; Ohta, H. Electron spin resonance studies of P and B codoped Si nanocrystals. Appl. Phys. Lett. 2008, 93 (2), 021920. (33) Marshall, A. R.; Beard, M. C.; Johnson, J. C. Nongeminate radiative recombination of free charges in cation-exchanged PbS quantum dot films. Chem. Phys. 2016, 471, 75−80. (34) Fujii, M.; Yamaguchi, Y.; Takase, Y.; Ninomiya, K.; Hayashi, S. Control of photoluminescence properties of Si nanocrystals by simultaneously doping n-and p-type impurities. Appl. Phys. Lett. 2004, 85 (7), 1158−1160.

N.X.C., R.L., and T.G. conceived the project, provided experimental interpretation, and wrote the manuscript. N.X.C. and R.L. conducted the induced absorption and photoluminescence measurements. N.X.C. performed the linear absorption and quantum yield measurements. A.L. prepared the cosputtered samples under supervision and facilitation of M.F., and C.d.W. guided the initial induced absorption measurements. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS The work was financially supported by the Technology Foundation STW, by NanoNextNL, by Stichting voor Fundamenteel Onderzoek der Materie (FOM), and R.L. acknowledges the National Renewable Energy Laboratory (NREL) LDRD Program for the award of the Nozik postdoctoral Fellowship to perform this research.



REFERENCES

(1) Schaller, R. D.; Klimov, V. I. High efficiency carrier multiplication in PbSe nanocrystals: implications for solar energy conversion. Phys. Rev. Lett. 2004, 92 (18), 186601. (2) Trinh, M. T.; Houtepen, A. J.; Schins, J. M.; Hanrath, T.; Piris, J.; Knulst, W.; et al. In spite of recent doubts carrier multiplication does occur in PbSe nanocrystals. Nano Lett. 2008, 8 (6), 1713−1718. (3) Luo, J. W.; Franceschetti, A.; Zunger, A. Carrier multiplication in semiconductor nanocrystals: theoretical screening of candidate materials based on band-structure effects. Nano Lett. 2008, 8 (10), 3174−3181. (4) Pijpers, J. J. H.; Ulbricht, R.; Tielrooij, K. J.; Osherov, A.; Golan, Y.; Delerue, C. Assessment of carrier-multiplication efficiency in bulk PbSe and PbS. Nat. Phys. 2009, 5 (11), 811. (5) Govoni, M.; Marri, I.; Ossicini, S. Carrier multiplication between interacting nanocrystals for fostering silicon-based photovoltaics. Nat. Photonics 2012, 6 (10), 672−679. (6) Nozik, A. J. Nanoscience and nanostructures for photovoltaics and solar fuels. Nano Lett. 2010, 10 (8), 2735−2741. (7) Wippermann, S.; Vörös, M.; Rocca, D.; Gali, A.; Zimanyi, G.; Galli, G. High-pressure core structures of Si nanoparticles for solar energy conversion. Phys. Rev. Lett. 2013, 110 (4), 046804. (8) Wippermann, S.; He, Y.; Vörös, M.; Galli, G. Novel silicon phases and nanostructures for solar energy conversion. Appl. Phys. Rev. 2016, 3 (4), 040807. (9) Schaller, R. D.; Pietryga, J. M.; Klimov, V. I. Carrier multiplication in InAs nanocrystal quantum dots with an onset defined by the energy conservation limit. Nano Lett. 2007, 7 (11), 3469−3476. (10) Stubbs, S. K.; Hardman, S. J.; Graham, D. M.; Spencer, B. F.; Flavell, W. R.; Glarvey, P.; et al. Efficient carrier multiplication in InP nanoparticles. Phys. Rev. B: Condens. Matter Mater. Phys. 2010, 81 (8), 081303. (11) Semonin, O. E.; Luther, J. M.; Choi, S.; Chen, H. Y.; Gao, J.; Nozik, A. J.; et al. Peak external photocurrent quantum efficiency exceeding 100% via MEG in a quantum dot solar cell. Science 2011, 334 (6062), 1530−1533. (12) Beard, M. C.; Knutsen, K. P.; Yu, P.; Luther, J. M.; Song, Q.; Metzger, W. K.; et al. Multiple exciton generation in colloidal silicon nanocrystals. Nano Lett. 2007, 7 (8), 2506−2512. (13) Saeed, S.; De Weerd, C.; Stallinga, P.; Spoor, F. C.; Houtepen, A. J.; Da Siebbeles, L.; et al. Carrier multiplication in germanium nanocrystals. Light: Sci. Appl. 2015, 4 (2), e251. (14) Yan, Y.; Crisp, R. W.; Gu, J.; Chernomordik, B. D.; Pach, G. F.; Marshall, A. R. Multiple exciton generation for photoelectrochemical hydrogen evolution reactions with quantum yields exceeding 100%. Nat. Energy 2017, 2, 17052. 2848

DOI: 10.1021/acsphotonics.8b00144 ACS Photonics 2018, 5, 2843−2849

Article

ACS Photonics (35) Ossicini, S.; Degoli, E.; Iori, F.; Luppi, E.; Magri, R.; Cantele, G.; et al. Simultaneously B-and P-doped silicon nanoclusters: Formation energies and electronic properties. Appl. Phys. Lett. 2005, 87 (17), 173120. (36) De Boer, W.; Timmerman, D.; Gregorkiewicz, T.; Zhang, H.; Buma, W.; Poddubny, A.; et al. Self-trapped exciton state in Si nanocrystals revealed by induced absorption. Phys. Rev. B: Condens. Matter Mater. Phys. 2012, 85 (16), 161409. (37) Limpens, R.; Gregorkiewicz, T. Spectroscopic investigations of dark Si nanocrystals in SiO2 and their role in external quantum efficiency quenching. J. Appl. Phys. 2013, 114 (7), 074304. (38) Chung, N. X.; Limpens, R.; Gregorkiewicz, T. Photoluminescence Quantum Yield in Ensembles of Si Nanocrystals. Adv. Opt. Mater. 2017, 5 (4), 1600709. (39) Valenta, J.; Greben, M.; Gutsch, S.; Hiller, D.; Zacharias, M. Effects of inter-nanocrystal distance on luminescence quantum yield in ensembles of Si nanocrystals. Appl. Phys. Lett. 2014, 105 (24), 243107. (40) Limpens, R.; Luxembourg, S. L.; Weeber, A. W.; Gregorkiewicz, T. Emission efficiency limit of Si nanocrystals. Sci. Rep. 2016, 6, 19566. (41) Iori, F.; Ossicini, S. Effects of simultaneous doping with boron and phosphorous on the structural, electronic and optical properties of silicon nanostructures. Phys. E 2009, 41 (6), 939−946. (42) Nomoto, K.; Sugimoto, H.; Breen, A.; Ceguerra, A. V.; Kanno, T.; Ringer, S. P.; et al. Atom Probe Tomography Analysis of Boron and/or Phosphorus Distribution in Doped Silicon Nanocrystals. J. Phys. Chem. C 2016, 120 (31), 17845−17852. (43) Aerts, M.; Bielewicz, T.; Klinke, C.; Grozema, F. C.; Houtepen, A. J.; Schins, J. M.; et al. Highly efficient carrier multiplication in PbS nanosheets. Nat. Commun. 2014, 5, 3789. (44) Hanna, M.; Nozik, A. Solar conversion efficiency of photovoltaic and photoelectrolysis cells with carrier multiplication absorbers. J. Appl. Phys. 2006, 100 (7), 074510. (45) Beard, M. C.; Midgett, A. G.; Hanna, M. C.; Luther, J. M.; Hughes, B. K.; Nozik, A. J. Comparing multiple exciton generation in quantum dots to impact ionization in bulk semiconductors: implications for enhancement of solar energy conversion. Nano Lett. 2010, 10 (8), 3019−3027. (46) Limpens, R.; Fujii, M.; Neale, N. R.; Gregorkiewicz, T. Negligible Electronic Interaction between Photoexcited Electron− Hole Pairs and Free Electrons in Phosphorus−Boron Co-Doped Silicon Nanocrystals. J. Phys. Chem. C 2018, 122, 6397. (47) Marri, I.; Govoni, M.; Ossicini, S. Carrier Multiplication in Silicon Nanocrystals: Theoretical Methodologies and Role of the Passivation. Phys. Stat. Sol. (C) 2017, 14 (12), na.

2849

DOI: 10.1021/acsphotonics.8b00144 ACS Photonics 2018, 5, 2843−2849