Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Tripled yield in direct-drive laser fusion through statistical modelling

Abstract

Focusing laser light onto a very small target can produce the conditions for laboratory-scale nuclear fusion of hydrogen isotopes. The lack of accurate predictive models, which are essential for the design of high-performance laser-fusion experiments, is a major obstacle to achieving thermonuclear ignition. Here we report a statistical approach that was used to design and quantitatively predict the results of implosions of solid deuterium–tritium targets carried out with the 30-kilojoule OMEGA laser system, leading to tripling of the fusion yield to its highest value so far for direct-drive laser fusion. When scaled to the laser energies of the National Ignition Facility (1.9 megajoules), these targets are predicted to produce a fusion energy output of about 500 kilojoules—several times larger than the fusion yields currently achieved at that facility. This approach could guide the exploration of the vast parameter space of thermonuclear ignition conditions and enhance our understanding of laser-fusion physics.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Target and laser pulse shape in the first phase of the Optimization Campaign.
Fig. 2: Prediction performance of the statistical model.
Fig. 3: Application of the statistical model to the experimental design of the DT target.
Fig. 4: Comparison of prediction accuracies of the statistical model and the simulated 1D yield.
Fig. 5: NIF-extrapolated fusion yields for selected OMEGA implosions.

Similar content being viewed by others

Data availability

Raw data were generated at the LLE’s OMEGA Laser Facility. Derived data supporting the findings of this study are available from the corresponding author upon reasonable request.

References

  1. Nuckolls, J., Wood, L., Thiessen, A. & Zimmerman, G. Laser compression of matter to super-high densities: thermonuclear (CTR) applications. Nature 239, 139–142 (1972).

    Article  ADS  CAS  Google Scholar 

  2. Atzeni, S. & Meyer-ter-Vehn, J. The Physics of Inertial Fusion: Beam Plasma Interaction, Hydrodynamics, Hot Dense Matter 1st edn (Oxford Univ. Press, Oxford, 2004).

    Book  Google Scholar 

  3. Slutz, S. A. et al. Pulsed-power-driven cylindrical liner implosions of laser preheated fuel magnetized with an axial field. Phys. Plasmas 17, 056303 (2010).

    Article  ADS  Google Scholar 

  4. Craxton, R. S. et al. Direct-drive inertial confinement fusion: a review. Phys. Plasmas 22, 110501 (2015).

    Article  ADS  Google Scholar 

  5. Lindl, J. D. Inertial Confinement Fusion: The Quest for Ignition and Energy Gain Using Indirect Drive (Springer-Verlag, New York, 1998).

    Google Scholar 

  6. Giorgini, S., Pitaevskii, L. P. & Stringari, S. Theory of ultracold atomic Fermi gases. Rev. Mod. Phys. 80, 1215–1274 (2008).

    Article  ADS  CAS  Google Scholar 

  7. Christopherson, A. R. et al. A comprehensive alpha-heating model for inertial confinement fusion. Phys. Plasmas 25, 012703 (2018).

    Article  ADS  Google Scholar 

  8. Betti, R. et al. Alpha heating and burning plasmas in inertial confinement fusion. Phys. Rev. Lett. 114, 255003 (2015).

    Article  ADS  CAS  Google Scholar 

  9. Hurricane, O. A. et al. Inertially confined fusion plasmas dominated by alpha-particle self-heating. Nat. Phys. 12, 800–806 (2016).

    Article  CAS  Google Scholar 

  10. Betti, R. et al. Thermonuclear ignition in inertial confinement fusion and comparison with magnetic confinement. Phys. Plasmas 17, 058102 (2010).

    Article  ADS  Google Scholar 

  11. Chang, P. Y. et al. Generalized measurable ignition criterion for inertial confinement fusion. Phys. Rev. Lett. 104, 135002 (2010).

    Article  ADS  CAS  Google Scholar 

  12. Spears, B. K. et al. Performance metrics for inertial confinement fusion implosions: aspects of the technical framework for measuring progress in the National Ignition Campaign. Phys. Plasmas 19, 056316 (2012).

    Article  ADS  Google Scholar 

  13. Betti, R. & Hurricane, O. A. Inertial-confinement fusion with lasers. Nat. Phys. 12, 435–448 (2016); corrigendum 12, 717 (2016).

    Article  CAS  Google Scholar 

  14. Le Pape, S. et al. Fusion energy output greater than the kinetic energy of an imploding shell at the National Ignition Facility. Phys. Rev. Lett. 120, 245003 (2018).

    Article  ADS  Google Scholar 

  15. Hurricane, O. A. et al. Fuel gain exceeding unity in an inertially confined fusion implosion. Nature 506, 343–348 (2014); corrigendum 510, 432 (2014).

    Article  ADS  CAS  Google Scholar 

  16. Campbell, E. M. & Hogan, W. J. The National Ignition Facility – applications for inertial fusion energy and high-energy-density science. Plasma Phys. Contr. Fusion 41, B39–B56 (1999).

    Article  CAS  Google Scholar 

  17. Edwards, M. J. et al. Progress towards ignition on the National Ignition Facility. Phys. Plasmas 20, 070501 (2013).

    Article  ADS  Google Scholar 

  18. Lindl, J. et al. Review of the National Ignition Campaign 2009–2012 Phys. Plasmas 21, 020501 (2014); erratum 21, 129902 (2014).

    Article  ADS  Google Scholar 

  19. Goncharov, V. N. et al. National direct-drive program on OMEGA and the National Ignition Facility. Plasma Phys. Contr. Fusion 59, 014008 (2017).

    Article  ADS  Google Scholar 

  20. Marinak, M. M. et al. Three-dimensional HYDRA simulations of National Ignition Facility targets. Phys. Plasmas 8, 2275–2280 (2001).

    Article  ADS  CAS  Google Scholar 

  21. Zimmerman, G., Kershaw, D., Bailey, D. & Harte, J. The LASNEX code for inertial confinement fusion. J. Opt. Soc. Am. 68, 549 (1978).

    ADS  Google Scholar 

  22. Radha, P. B. et al. Two-dimensional simulations of plastic-shell, direct-drive implosions on OMEGA. Phys. Plasmas 12, 032702 (2005).

    Article  ADS  Google Scholar 

  23. Delettrez, J., Epstein, R., Richardson, M. C., Jaanimagi, P. A. & Henke, B. L. Effect of laser illumination nonuniformity on the analysis of time-resolved X-ray measurements in UV spherical transport experiments. Phys. Rev. A 36, 3926–3934 (1987).

    Article  ADS  CAS  Google Scholar 

  24. Atzeni, S. et al. Fluid and kinetic simulation of inertial confinement fusion plasmas. Comput. Phys. Commun. 169, 153–159 (2005).

    Article  ADS  CAS  Google Scholar 

  25. Clark, D. S. et al. Radiation hydrodynamics modeling of the highest compression inertial confinement fusion ignition experiment from the National Ignition Campaign. Phys. Plasmas 22, 022703 (2015).

    Article  ADS  Google Scholar 

  26. Igumenshchev, I. V. et al. Three-dimensional modeling of direct-drive cryogenic implosions on OMEGA. Phys. Plasmas 23, 052702 (2016).

    Article  ADS  Google Scholar 

  27. Igumenshchev, I. V. et al. Crossed-beam energy transfer in implosion experiments OMEGA. Phys. Plasmas 17, 122708 (2010).

    Article  ADS  Google Scholar 

  28. Michel, D. T. et al. Measurements of the ablation-front trajectory and low-mode nonuniformity in direct-drive implosions using X-ray self-emission shadowgraphy. High Power Laser Sci. Eng. 3, e19 (2015).

    Article  Google Scholar 

  29. Goncharov, V. N. et al. Early stage of implosion in inertial confinement fusion: shock timing and perturbation evolution. Phys. Plasmas 13, 012702 (2006).

    Article  ADS  Google Scholar 

  30. Hu, S. X. et al. First-principles equation of state of polystyrene and its effect on inertial confinement fusion implosions. Phys. Rev. E 92, 043104 (2015).

    Article  ADS  CAS  Google Scholar 

  31. Boehly, T. R. et al. Initial performance results of the OMEGA Laser System. Opt. Commun. 133, 495–506 (1997).

    Article  ADS  CAS  Google Scholar 

  32. Regan, S. P. et al. Demonstration of fuel hot-spot pressure in excess of 50 Gbar for direct-drive, layered deuterium–tritium implosions on OMEGA. Phys. Rev. Lett. 117, 025001 (2016).

    Article  ADS  CAS  Google Scholar 

  33. Bose, A. et al. Core conditions for alpha heating attained in direct-drive inertial confinement fusion. Phys. Rev. E 94, 011201 (2016).

    Article  ADS  CAS  Google Scholar 

  34. Robey, H. F. et al. Measurement of high-pressure shock waves in cryogenic deuterium-tritium ice layered capsule implosions on NIF. Phys. Rev. Lett. 111, 065003 (2013).

    Article  ADS  CAS  Google Scholar 

  35. Betti, R. et al. Shock ignition of thermonuclear fuel with high areal density. Phys. Rev. Lett. 98, 155001 (2007).

    Article  ADS  CAS  Google Scholar 

  36. Kilkenny, J. D. et al. A review of the ablative stabilization of the Rayleigh–Taylor instability in regimes relevant to inertial confinement fusion. Phys. Plasmas 1, 1379–1389 (1994).

    Article  ADS  CAS  Google Scholar 

  37. Brouillette, M. The Richtmyer–Meshkov instability. Annu. Rev. Fluid Mech. 34, 445–468 (2002).

    Article  ADS  MathSciNet  Google Scholar 

  38. Peterson, J. L. et al. Enhancing hohlraum design with artificial neural networks. Bull. Am. Phys. Soc. 62 http://meetings.aps.org/link/BAPS.2017.DPP.BO7.6 (2017).

  39. Salvatier, J., Wiecki, T. V. & Fonnesbeck, C. Probabilistic programming in python using PyMC3. PeerJ Comput. Sci. 2, e55 (2016).

    Article  Google Scholar 

  40. Boehly, T. R. et al. Optical and plasma smoothing of laser imprinting in targets driven by lasers with SSD bandwidths up to 1 THz. Phys. Plasmas 8, 2331–2337 (2001).

    Article  ADS  CAS  Google Scholar 

  41. Hu, S. X. et al. Understanding the effects of laser imprint on plastic-target implosions on OMEGA. Phys. Plasmas 23, 102701 (2016).

    Article  ADS  Google Scholar 

  42. Murphy, T. J., Chrien, R. E. & Klare, K. A. Interpretation of neutron time-of-flight signals from current-mode detectors. Rev. Sci. Instrum. 68, 610–613 (1997).

    Article  ADS  CAS  Google Scholar 

  43. Murphy, T. J. The effect of turbulent kinetic energy on inferred ion temperature from neutron spectra. Phys. Plasmas 21, 072701 (2014).

    Article  ADS  Google Scholar 

  44. Woo, K. M. et al. Effects of residual kinetic energy on yield degradation and ion temperature asymmetries in inertial confinement fusion implosions. Phys. Plasmas 25, 052704 (2018).

    Article  ADS  Google Scholar 

  45. Nora, R. et al. Theory of hydro-equivalent ignition for inertial fusion and its applications to OMEGA and the National Ignition Facility. Phys. Plasmas 21, 056316 (2014).

    Article  ADS  Google Scholar 

  46. Bose, A. et al. Analysis of trends in implosion observables for direct-drive cryogenic implosions on OMEGA. Phys. Plasmas 25, 062701 (2018).

    Article  ADS  Google Scholar 

  47. Smalyuk, V. A. et al. Role of hot-electron preheating in the compression of direct-drive imploding targets with cryogenic D2 ablators. Phys. Rev. Lett. 100, 185005 (2008).

    Article  ADS  CAS  Google Scholar 

  48. Forrest, C. J. et al. High-resolution spectroscopy used to measure inertial confinement fusion neutron spectra on OMEGA. Rev. Sci. Instrum. 83, 10D919 (2012).

    Article  CAS  Google Scholar 

  49. Casey, D. T. et al. The magnetic recoil spectrometer for measurements of the absolute neutron spectrum at OMEGA and the NIF. Rev. Sci. Instrum. 84, 043506 (2013).

    Article  ADS  CAS  Google Scholar 

  50. Donaldson, W. R., Boni, R., Keck, R. L. & Jaanimagi, P. A. A self-calibrating, multichannel streak camera for inertial confinement fusion applications. Rev. Sci. Instrum. 73, 2606–2615 (2002).

    Article  ADS  CAS  Google Scholar 

  51. Edgell, D. H. et al. Three-dimensional characterization of cryogenic target ice layers using multiple shadowgraph views. Fus. Sci. Technol. 49, 616–625 (2006).

    Article  CAS  Google Scholar 

  52. Hu, S. X., Militzer, B., Goncharov, V. N. & Skupsky, S. First-principles equation-of-state table of deuterium for inertial confinement fusion applications. Phys. Rev. B. 84, 224109 (2011).

    Article  ADS  Google Scholar 

  53. Zhou, C. D. & Betti, R. A measurable Lawson criterion and hydro-equivalent curves for inertial confinement fusion. Phys. Plasmas 15, 102707 (2008); correction 16, 079905 (2009).

    Article  ADS  Google Scholar 

  54. Hoffman, M. D. & Gelman, A. The no-U-turn sampler: adaptively setting path lengths in Hamiltonian Monte Carlo. J. Mach. Learn. Res. 15, 1593–1623 (2014).

    MathSciNet  MATH  Google Scholar 

  55. Goncharov, V. N. et al. Improving the hot-spot pressure and demonstrating ignition hydrodynamic equivalence in cryogenic deuterium–tritium implosions on OMEGA. Phys. Plasmas 21, 056315 (2014).

    Article  ADS  Google Scholar 

  56. Marozas, J. A. et al. First observation of cross-beam energy transfer mitigation for direct-drive inertial confinement fusion implosions using wavelength detuning at the National Ignition Facility. Phys. Rev. Lett. 120, 085001 (2018).

    Article  ADS  CAS  Google Scholar 

  57. Rosenberg, M. J. et al. Origins and scaling of hot-electron preheat in ignition-scale direct-drive inertial confinement fusion experiments. Phys. Rev. Lett. 120, 055001 (2018).

    Article  ADS  CAS  Google Scholar 

  58. Zhou, C. D. & Betti, R. Hydrodynamic relations for direct-drive fast-ignition and conventional inertial confinement fusion implosions. Phys. Plasmas 14, 072703 (2007).

    Article  ADS  Google Scholar 

Download references

Acknowledgements

This material is based on work supported by the Department of Energy National Nuclear Security Administration under award numbers DE-NA0003856 and DENA0001944, the Department of Energy Office of Fusion Energy Sciences under award number DE-FC02-04ER54789, the University of Rochester and the New York State Energy Research and Development Authority. This report was prepared as an account of work sponsored by an agency of the US Government. Neither the US Government nor any agency thereof, nor any of their employees, makes any warranty, express or implied, or assumes any legal liability or responsibility for the accuracy, completeness or usefulness of any information, apparatus, product or process disclosed, or represents that its use would not infringe privately owned rights. Reference herein to any specific commercial product, process or service by trade name, trademark, manufacturer or otherwise does not necessarily constitute or imply its endorsement, recommendation or favouring by the US Government or any agency thereof. The views and opinions of authors expressed herein do not necessarily state or reflect those of the US Government or any agency thereof.

Reviewer information

Nature thanks I. Kaganovich, R. Scott and the other anonymous reviewer(s) for their contribution to the peer review of this work.

Author information

Authors and Affiliations

Authors

Contributions

All authors contributed to the work presented in this paper. A.A.S., A.M. and J. Palastro were responsible for theoretical and experimental work on hot-electron pre-heating, and J. Peebles investigated high--mode instabilities due to laser imprinting, both of which were used in constraining the optimization process of this work.

Corresponding author

Correspondence to V. Gopalaswamy.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Extended data figures and tables

Extended Data Fig. 1 Posterior predictive distributions for the parameters in equation (6).

The intercept parameter refers to the constant Cj in equation (5), ε refers to the noise parameter in equation (8) and the remaining parameters are the μi values for the corresponding simulated quantities, as in equation (5). The Markov chains for each parameter are converged over 5,000 steps. The tails of the distributions have the same sign as the mean and are not excessively wide when compared to the mean. All these indicate that the model is well specified.

Extended Data Fig. 2 Posterior predictive distributions for the parameters in equation (7).

The intercept parameter refers to the constant Cj in equation (5), ε refers to the noise parameter in equation (8) and the remaining parameters are the μi values for the corresponding simulated quantities, as in equation (5). As in Extended Data Fig. 1, these results indicate that the model is well specified; we note that this does not guarantee that the model will have any predictive power.

Extended Data Fig. 3 The effect of RT on the yield is less than 10% for the vast majority of implosions.

Vertical error bars are one standard deviation of \({R}_{T}^{-0.6}\). Because the yield is proportional to \({R}_{T}^{-0.6}\) in equation (7), and the vast majority of implosions carried out at OMEGA are repeatable, the empirical correction factor in equation (7) varies by less than 10% for most OMEGA implosions. This justifies the use of this parameter to account for the occasional shot with large random nonuniformities.

Extended Data Fig. 4 Areal densities for high-convergence implosions are well predicted.

The dashed line is the y = x line on which the data points would lie if the experimental areal density were perfectly modelled by the power-law dependence. As the convergence increases, the safety margin for a nominal implosion becomes thinner and makes designing these implosions more challenging. A predictive model for the areal density was built using the framework presented in this paper (training data are shown by blue circles). The areal density was increased from that of shot 87266 (green circle) over four shot days (orange circles) until the highest values were reached in July 2018. The yields for these implosions are marked.

Extended Data Fig. 5 Comparison of prediction accuracies of the statistical model and 1D simulations.

a, b, Predictions for the areal density (a) and neutron yield (b) including the second phase of the Optimization Campaign. The predictions from the statistical model (blue) remain accurate for implosions from both phases of the Optimization Campaign, whereas the corresponding simulated 1D quantities from LILAC (red) remain inaccurate and overpredict potential increases in performance. Horizontal error bars and centre values represent one standard deviation and mean, respectively, for 500 draws from the posterior distribution of the statistical model. Vertical error bars represent one standard deviation for the neutron yield and areal density detectors. The dashed line is the y = x line on which the data points would lie if the prediction perfectly matched the measurement.

Extended Data Fig. 6 Power-law dependence of the simulated yield on simulated parameters for initial conditions used in the Optimization Campaign.

The dashed line is the y = x line on which the data points would lie if the simulated yield were perfectly modelled by the power-law dependence. The simulated yield (Ysim) is well represented by power-law relations to the simulated implosion velocity (\({V}_{{\rm{i}}{\rm{m}}{\rm{p}}}^{{\rm{s}}{\rm{i}}{\rm{m}}}\)), mass (\({M}_{{\rm{s}}{\rm{t}}{\rm{a}}{\rm{g}}}^{{\rm{s}}{\rm{i}}{\rm{m}}}\)) and areal density (ρRsim). Although the exponents are not identical to those from analytical theory, physical intuition has provided a good basis for variable selection.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Gopalaswamy, V., Betti, R., Knauer, J.P. et al. Tripled yield in direct-drive laser fusion through statistical modelling. Nature 565, 581–586 (2019). https://doi.org/10.1038/s41586-019-0877-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-019-0877-0

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing AI and Robotics

Sign up for the Nature Briefing: AI and Robotics newsletter — what matters in AI and robotics research, free to your inbox weekly.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing: AI and Robotics