Burning plasma achieved in inertial fusion – Nature.com

Terminology

Definitions of commonly used mathematical symbols are summarized in Extended Data Table 2.

Reproducibility

Experiments subsequent to those described in this Article have demonstrated the reproducibility of a burning-plasma state at NIF, with two additional experiments that have performance comparable to the highest-performing experiments in this Article. These newer experiments, N210307 and N210605, were conducted in the Hybrid E platform. N210307 repeated N210207, albeit using a capsule from a different fabrication batch and produced a yield of approximately 145 kJ with an experimentally inferred Qα = 1.34 ± 0.07 from the Hurricane criterion. Experiment N210605 reduced the thickness of the ice layer relative to N210207 and resulted in a lower yield (135 kJ) but high ion temperature, with Qα = 1.40 ± 0.10, again from the Hurricane criterion. These additional experiments confirm that the burning-plasma state is reproducible at NIF, and full details and analysis on them will be presented in future publications.

Inferred hot-spot conditions

Hot-spot conditions must be inferred from measured quantities using a model. The simplest hot-spot model is to assume an isobaric volume of uniform conditions, as used in a previous work14 between equations 2 and 3, in which case the hot-spot number density is given by

$$n=1.2times {10}^{6}sqrt{frac{Y}{langle sigma vrangle {V}_{{rm{hs}}}tau }},$$

(2)

where Y is the fusion yield in J, σν is the fusion reactivity, which depends on the ion temperature (Ti), Vhs is the hot-spot volume in cm3, and τ is the burn duration in s, for equimolar DT mixtures. The remaining hot-spot quantities follow from the inferred density, including the pressure (P = (1 + Z)nkBTi, with kB Boltzmann’s constant), hot-spot energy (Ehs = 1.5PVhs), and areal density (({rho R}_{{rm{h}}{rm{s}}}=(2.5n/{N}_{{rm{a}}})sqrt[3]{3{V}_{{rm{h}}{rm{s}}}/4pi })).

A more detailed inference is to use a one-dimensional (1D) profile in radius for temperature and density, maintaining the isobaric assumption. A conduction-limited profile follows the expression45:

$$T(r)={T}_{min }+({T}_{0}-{T}_{min }){left[1-{left(frac{r}{{R}_{0}}right)}^{2}right]}^{frac{1}{1+beta }},$$

(3)

where Tmin is the temperature at the boundary, T0 is the central temperature and R0 is the hot-spot boundary. β is the thermal conductivity power law, 2.5 from classical Spitzer conductivity. Following a previous work26 we use a lower value, β = 2/3, which accounts for additional physics, dynamical processes and reproduces radiation–hydrodynamics simulations. The density profile is then determined by the isobaric assumption through PnT being constant. Tmin is taken as 1 keV leaving T0, R0 and P as free parameters in the model; the data are compared to synthetic data calculated from this 1D profile with the model parameters adjusted to minimize residuals. As in the zero-dimensional (0D) model, the hot-spot energy simply follows from pressure and volume, and the areal density is the mass density integrated over the inferred radial profile.

In either dimensionality the model’s radius is matched to the experimental measurements, which take a contour of emission level, by calculating synthetic emission images to calculate an equivalent contour radius. The measurements include 2D and three-dimensional (3D) asymmetries, so an equivalent spherical volume, and radius, are calculated using the modal decompositions, where the emission contour measured from the equator (Req) and pole (Rpo) are

$${R}_{{rm{eq}}}={P}_{0}{1+sum _{{ell }}{rm{delta }}{P}_{{ell }}times {P}_{{ell }}(cos ,theta )},$$

(4)

$${R}_{{rm{po}}}={M}_{0}{1+sum _{m}{rm{delta }}{M}_{m}times ,cos ,[m(varphi -{varphi }_{i})]},$$

(5)

$${R}_{{rm{hs}}}=frac{{R}_{{rm{eq}}}{R}_{{rm{po}}}}{{M}_{0}},$$

(6)

where P0 and M0 are the average measured sizes from each view, δPMm) is the relative modal amplitude, often referred to as P/P0 with the nomenclature above chosen for clarity. P are the Legendre polynomials, and ϕm are the orientation of the azimuthal modes. Rhs is the hot-spot radius as a function of both θ and ϕ, which is integrated to obtain the volume. Here, the dominant parameters are P0, M0 and δP2, which are given in ref. 7.

Implosion velocity (vimp) is inferred using a rocket model of the implosion46 constrained by both supporting experiments, especially in-flight radiography, and the measured time of peak nuclear production on each experiment. The inferred yield amplification given in Extended Data Table 1 is a function of the measured yield, shell compression and fuel mass (mfuel); both the velocity and Yamp inferences use the prescription given in a previous work26. The fuel kinetic energy then follows from (frac{1}{2}{m}_{{rm{fuel}}}{v}_{{rm{imp}}}^{2}). Our techniques for inferring the PdV work done on the fuel are discussed in the following section.

A comparison of inferred values using 0D and 1D models are shown in Extended Data Table 3. Inferred pressures are highly consistent between these calculations, whereas hot-spot energies and areal densities are higher in the 1D model owing to substantial mass near the 1 keV temperature cut-off.

Inferring G
fuel

The total fusion yield produced by a mass (m) of DT, over a characteristic confinement time, τ, is Y ≈ 5mPατ—with Pα = 8.2 × 1024ρσν in GJ g−1 s−1 the specific DT fusion power for a given mass density, ρ, of DT with reaction rate σν—and the internal energy in that DT is Ehs = cDTmTi. Therefore, one can write (O.A.H. et al., manuscript in preparation)

$${G}_{{rm{fuel}}}=frac{Y}{{E}_{P{rm{d}}V,{rm{tot}}}}approx frac{frac{Y}{{E}_{{rm{hs}}}}}{1+frac{{E}_{{rm{fuel}}}}{{E}_{{rm{hs}}}}-frac{q}{10}frac{Y}{{E}_{{rm{hs}}}}},$$

(7)

with

$$frac{Y}{{E}_{{rm{hs}}}}approx 4.6times {10}^{26}Pfrac{langle sigma vrangle }{{T}^{2}}tau ,$$

(8)

where P is in Gbar, Ti in keV, and τ in s. In equation (7), the total energy delivered by PdV work, EPdV,tot, is determined from the hot spot and compressed, but cold, DT fuel energy at stagnation, Ehs and Efuel, respectively, at peak compression. The last term in the denominator represents a correction for additional energy retained by self-heating of the fuel from α-particle deposition but not then lost as bremsstrahlung. So, EPdV,tot ≈ Ehs + Efuel − qY/10, where q is a ‘quality’ factor, 0 ≤ q ≤ 1, measuring the ability of the implosion to retain self-heating energy (O.A.H. et al., manuscript in preparation). Here we use q ≈ 0.7, inferred from simulations, and the factor of 10 results from one-fifth of the fusion energy released as α-particles and half of those produced up until the time of peak fusion burn. Albeit generally arrived at in a different fashion than above, the product P(σν/T2)τ is Lawson’s9 parameter for ignition. Figure 2a uses the useful reaction-rate approximation (langle sigma vrangle approx 4.2times {10}^{-20}{T}_{{rm{i}}}^{3.6}) (in units of cm3 s−1 for ion temperature range 3.5 < Ti < 6.5 keV) to simplify the abscissa.

An expression for the fuel gain is given in equation (7). The yield is measured and the hot-spot energy is inferred as described in the previous section. Precisely determining the cold-fuel energy from data is not straightforward. For the purposes of this analysis we actually require the total PdV work done on all the DT. This is at a minimum the fuel kinetic energy and internal energy at peak velocity, which are both inferred. This neglects any work done by the inflowing remaining ablator material on the fuel, which can occur in these implosions. In this case the hot-spot energy is more than half the previous estimate; in this scenario we assume equipartition between the hot spot and cold fuel to evaluate equation (7).

Inferred PdV work

The primary uncertainty in the Betti metric8 is in the inference of PdV work on the hot spot. Here we use three methodologies: two inferences using an analytic model, and a direct extraction of PdV work from simulations that match the experimental observables.

We use the hydrodynamic piston model of an implosion described previously34. This analytic model abstracts the implosion process using opposed pistons to represent the imploding shell. In spherical geometry, the stagnation pressure from this mechanical work on the hot spot is given by (equation 24in ref. 34):

$${P}_{{rm{piston}}}=frac{rho {rm{delta }}{R}_{{rm{ave}}}{v}_{{rm{imp}}}^{2}}{{R}_{{rm{hs}}}}(1-{f}^{2},),$$

(9)

where ρδRave is the average shell areal density, calculated from the measured neutron ‘down-scattered ratio’ (DSR) using the relation ρδRave ≈ 19.3DSR, vimp is the implosion velocity and Rhs is the average hot-spot radius (which can be obtained from the volume, Vhs, given in Extended Data Table 1). The factor f2 represents the effect of mode-1 asymmetry and is a measure of the residual kinetic energy (kinetic energy that is never converted into internal energy) in the implosion.

From the piston pressure we obtain the hot-spot internal energy (Ehs) from

$${E}_{{rm{hs}}}=frac{3}{2}{P}_{{rm{piston}}}{V}_{{rm{hs}}}.$$

(10)

In the absence of α-heating (which adds energy to the hot-spot) and radiative X-ray losses, or when α-heating exactly balances X-ray losses, then Ehs = EPdV,hs. For low yield amplification implosions (Yamp < 1.5), X-ray losses dominate over α-heating energy gains, so Ehs < EPdV,hs. For higher yield amplification implosions (Yamp > 2), α-heating energy gains start to dominate over X-ray losses, so Ehs > EPdV,hs. The estimated values for these four experiments are given in Extended Data Table 4 as the piston methodology.

We can also estimate the stagnated fuel mass in a similar fashion, using

$${m}_{{rm{shell}}}=4{rm{pi }}{R}_{{rm{hs}}}^{2},rho {rm{delta }}{R}_{{rm{ave}}},$$

(11)

which allows us to then estimate the total mass that stagnates from mshell + mhs, with mhs from the hot-spot inferences described earlier. We then estimate the PdV work from

$${E}_{P{rm{d}}V,{rm{hs}}}=0.73{{rm{KE}}}_{{rm{fuel}}}frac{{m}_{{rm{shell}}}+{m}_{{rm{hs}}}}{{m}_{{rm{fuel}}}},$$

(12)

where mfuel is the initial fuel mass. The factor of 0.73 is derived from 1D simulations in which the imploding mass stagnates efficiently, and we drop the residual kinetic energy factor f2 because the inferred shell mass does not include non-stagnated material. This estimate leads to smaller estimates of EPdV,hs than the first empirical estimate, and are given in Extended Data Table 4 as the stagnated mass estimate.

For analysis of previously published campaigns we use the simple relation EPdV,hs ≈ (0.5–0.7)KEfuel(1 − f 2), this is easy to evaluate with the available data and the factor 0.5–0.7 accounts for a wide range of 1D to 2D/3D behaviour observed on past experiments. For comparison, the proportionality constant inferred from the first methodology (equation (10)) is between 0.60 and 0.73 for our four experiments.

We also use radiation–hydrodynamics simulations to estimate the PdV work done on these implosions. The first simulation-based methodology is to use 2D simulations with degradation mechanisms that match the observed performance, and interrogate the work done upon the mass elements that form the hot spot to infer EPdV,hs. The simulation methodology is described in ref. 6, and the values of EPdV,hs for this method are given in Extended Data Table 4. The same fusion performance can be generated with varying application of degradation mechanisms that either degrade EPdV,hs or do not; an estimate of the 2D simulation uncertainty of ±0.5 kJ is estimated by studying multiple simulations.

A similar energy-balance analysis can be done with 1D simulations, in which the work done upon the hot spot is well defined with a Lagrangian mesh. The 1D simulations are tuned to match the measured yields, but are expected to underestimate EPdV,hs since they cannot properly incorporate residual kinetic energy. This estimate is given in Extended Data Table 4 as an upper bound.

We have thus develop four methodologies for estimating EPdV,hs. In the main analysis we use a combination of the empirical piston model estimate as the more pessimistic data-based inference, and use the 2D simulated EPdV,hs as the most robust computational description of the experiments.

Modified Hurricane metric

At peak burn, the time rate of change of hot-spot volume, dV/dt, is nearly zero, and therefore so is the heating rate, so time integration is needed. Mathematically, a statement of a burning plasma appropriate for ICF is

$${int }_{0}^{{t}_{{rm{pf}}}}{P}_{{rm{alpha }}}{rm{d}}t > -{int }_{0}^{{t}_{min V}}frac{P}{m}{rm{d}}V,$$

(13)

where tpf is the time of peak fusion rate, and tminV is the time of minimum hot-spot volume.

The integrals in equation (13) are easily approximated2 without knowing the details of the actual implosion using the mathematical method of steepest descent; assuming that the thermodynamic quantities of interest, such as T, P, ρ, and so on, are impulsive, being highly peaked around the time of stagnation. Ultimately, the solution to equation (13), in terms of only burn-average hot-spot areal density, ρRhs, Ti and vimp is equation (1) after a correction to the original derivation.

A recent note from our colleagues at Los Alamos47 discovered an arithmetic error in the derivation of the criteria as published in ref. 3. The error is in going from equation 8 to equation 9 in ref. 3, in which the conversion to peak temperature (T0) to burn-averaged temperature (Ths) should be, for n ≈ 4,

$$begin{array}{c}{frac{langle sigma vrangle }{{T}_{0}}|}_{{T}_{0}}approx {left(frac{n+1}{n}right)}^{frac{n-1}{2}}{frac{langle sigma vrangle }{T}|}_{{T}_{{rm{hs}}}}\ ,approx ,1.40{frac{langle sigma vrangle }{T}|}_{{T}_{{rm{hs}}}},.end{array}$$

(14)

Additionally, we now believe that the inclusion of the fraction of α-particles stopping in the hot spot (fα) in ref. 3 was inappropriate. When considering the temperature evolution of a defined mass—for example, the self-heating criterion in equation (17)—this is necessary because fα is fundamentally the fraction of α-particle energy deposited into that mass. On the other hand, the burning-plasma criteria is one on the energy of the hot spot,

$${E}_{{rm{hs}}}={c}_{{rm{DT}}}{m}_{{rm{hs}}}{T}_{{rm{hs}}},$$

(15)

and α-particles that escape the hot spot still contribute to its energy via generation of additional hot-spot mass, as seen by examining the time derivative of the above:

$$begin{array}{ccc}frac{{rm{d}}{E}_{{rm{hs}}}}{{rm{d}}t} & = & {c}_{{rm{DT}}}left({m}_{{rm{hs}}}frac{{rm{d}}{T}_{{rm{hs}}}}{{rm{d}}t}+frac{{rm{d}}{m}_{{rm{hs}}}}{{rm{d}}t}{T}_{{rm{hs}}}right)\ & = & {m}_{{rm{hs}}},{f}_{{rm{alpha }}}{Q}_{{rm{alpha }}}+{m}_{{rm{hs}}}(1-{f}_{{rm{alpha }}}){Q}_{{rm{alpha }}}\ & = & {m}_{{rm{hs}}}{Q}_{{rm{alpha }}}.end{array}$$

(16)

Therefore, the inclusion of fα in a burning-plasma criterion is inappropriate. We note that not including an fα factor is consistent with other criteria, for example, ref. 8. With these two modifications to the criterion published in ref. 3 we use a new criterion (equation (1)). This modified criterion is slightly more restrictive for the burning-plasma threshold in the regime relevant to these experiments.

Model uncertainties for Hurricane’s metric

The Hurricane metric3 depends on more quantities than the Betti metric, although these quantities are more straightforward to infer than EPdV,hs. The metric reduces to equation (1) where ρRhs and vimp are inferred as described previously, and Ti is measured. σν contains some systematic uncertainty from the evaluation used. Data uncertainties are well defined for Ti and in the inference of ρRhs and vimp, and are propagated as described in the next section; the inferred ρRhs can also vary between models, which will be discussed.

Equation (1) depends on the fusion reactivity; in this work we use the evaluation from Bosch and Hale32. Recent publications have presented alternative evaluations48 which differ by about 2%. We note that the inferred (rho Rpropto 1/sqrt{langle sigma vrangle }) from equation (2), so the condition in equation (1) depends on the reactivity as (1/sqrt{langle sigma vrangle }). fα is also weakly increasing with ρR, leading to the condition being slightly less than square-root dependent on σν, so this criterion has <1% uncertainty from the choice of σν evaluation.

The Hurricane criterion is sensitive to the inferred hot-spot ρR, which can vary between models depending on the spatial dependence of ρ. As shown in Extended Data Table 3, the 0D and 1D hot-spot models agree quite well. We also check these values using a 3D reconstruction of the hot-spot density and temperature profiles (a yet unpublished method of L. Divol, but briefly described in ref. 35): for N201101 this gives a value of ρRhs ≈ 0.36–0.38 g cm−2 to the 1-keV contour for N201101 and ρRhs ≈ 0.35–0.36 g cm−2 for N201122. These values are consistent with the simple models described earlier.

Self-heating regime

The hot-spot per unit mass power balance is:

$${c}_{{rm{DT}}}frac{{rm{d}}T}{{rm{d}}t}={f}_{{rm{alpha }}}{P}_{{rm{alpha }}}-{f}_{{rm{b}}}{P}_{{rm{b}}}-{P}_{{rm{e}}}-frac{P}{m}frac{{rm{d}}V}{{rm{d}}t},$$

(17)

which describes the temporal evolution of the temperature (T) in terms of the balance of self heating (Pα) versus bremsstrahlung (Pb) and electron conduction (Pe) losses plus PdV work. Here electron conduction losses are calculated relative to a hot-spot boundary that is defined relative to a fraction of the peak burn rate or a specified ion temperature. Thermal conduction cools the hot spot while increasing the mass of the hot spot. Because the fusion burn rate is more strongly dependent on the temperature of the spot than its mass in the temperature range achieved by compression alone, α-heating must provide sufficient heating for the hot-spot temperature to increase in the presence of this conduction into an increasing mass. Hot-spot volume change, dV/dt, is negative on implosion, increasing T. During expansion the PdV term becomes an energy loss term. The bremsstrahlung loss can be enhanced beyond the emission of clean DT by the presence of high-Z contamination of the DT (that is, mix), by a fraction fb. In equation (17), fα is the fraction of α-particles stopped in the hot spot, evaluated using fits with modern stopping-power theory36.

Uncertainty analysis

We perform uncertainty analysis for all hot-spot quantities by propagating the normally distributed uncertainties in measured quantities through the 0D and 1D models described earlier. The model input parameters are those that fully describe the system, and are constrained by the measured yield, ion temperature, burn widths (from both X-rays and γ-rays), and volume from the 17% contour of neutron emissivity. Distributions of model parameters are generated using Markov chain Monte Carlo (MCMC), calculated with the tensorflow49 probability package. The log-likelihood function for MCMC is defined by the measurements and calculated with the log-likelihood function

$$-frac{1}{2}sum _{i}{left(frac{{m}_{i}-{y}_{i}}{{rm{delta }}{y}_{i}}right)}^{2},$$

(18)

which is summed over all observables (i) where mi is the model value, yi is the measured value and δyi is the uncertainty in the measurement. This methodology produces full distributions of the model parameters including any correlations, from the model parameter distributions we generate full distributions of all hot-spot parameters, some of which exhibit correlation, such as in the temperature and areal density required to evaluate the Hurricane metric, which are partially anti-correlated (evident in Fig. 3a). Other inferences, such as the implosion velocity or kinetic energy, are treated with normally distributed uncertainties that are uncorrelated with the hot-spot inferences.

Power-balance relations

In evaluating the power-balance relations relevant to equation (17) we use the following expressions for the individual terms:

$${P}_{{rm{alpha }}}=8.2times {10}^{24}rho langle sigma vrangle ,$$

(19)

$${P}_{{rm{b}}}=3.1times {10}^{7}rho sqrt{T},$$

(20)

$${P}_{{rm{e}}}=5.9times {10}^{3}frac{{T}^{3.5}}{rho {R}^{2}}.$$

(21)

In these expressions the specific powers are given in units of GJ g−1 s−1 and thus are multiplied by the inferred hot-spot mass to obtain power. ρ is the hot-spot mass in g cm−3, σν is the fusion reactivity evaluated as a function of temperature in cm3 s−1, T is the temperature in keV, and ρR is the hot-spot areal density in g cm−2. The self-heating power Pα is multiplied by the fraction of α-particle energy deposited in the hot spot (fα) using the evaluation published in ref. 36; for all four experiments, fα ≈ 0.77–0.80.

Leave a comment

Your email address will not be published. Required fields are marked *