Skip to content

First Wall

The first wall is the wall surface located closest to the core plasma in a fusion reactor. It directly receives radiation and particles from the plasma, protects the blanket, and plays a crucial role in recovering thermal energy. As one of the most challenging technical issues for practical fusion reactors, research and development spanning over half a century continues in this interdisciplinary field combining materials science, thermal engineering, and plasma physics.

The concept of the first wall dates back to the dawn of fusion research in the 1950s. Initially, it was expected that plasma-wall interactions could be minimized if plasma could be completely confined by magnetic fields, but as experiments progressed, the importance of plasma-wall interactions became recognized.

From the 1960s to the 1970s, as improved confinement performance achieved in Soviet tokamak experiments provided prospects for maintaining high-temperature plasma for extended periods, first wall design emerged as a serious engineering challenge. In particular, the plasma temperatures exceeding 10 million degrees achieved by the T-3 tokamak in 1968 made heat load issues on wall materials a reality.

In the 1980s, large tokamaks such as JET (Europe), TFTR (USA), and JT-60 (Japan) were constructed, and methodologies for first wall material selection and design were established. Carbon-based materials were widely used, but tritium retention issues became apparent, leading to a shift toward metallic materials (beryllium, tungsten) from the 1990s onward.

Through the ITER design process (1988-present), engineering design of the first wall has advanced dramatically, and design methodologies, material selection criteria, and testing/verification methods for plasma-facing components have been systematized.

The first wall is installed along the magnetic surfaces of the scrape-off layer and has the following functions:

  • First to receive neutron energy (neutrons are not confined by magnetic fields and scatter in all directions)
  • Receives electromagnetic radiation from plasma (bremsstrahlung, synchrotron radiation, etc.)
  • Receives neutral particles generated by charge exchange reactions
  • As the plasma-side surface of the blanket, minimizes impact on tritium breeding ratio
  • Suppresses the amount of impurities entering the plasma

Approximately 20% of fusion output (alpha particle energy) ultimately flows into the first wall and divertor. Of this, line radiation from impurities in the plasma and bremsstrahlung from plasma electrons are mainly transported to the first wall surface.

The main requirements for first wall design are summarized below.

CategoryRequirementTypical Specification
Thermal PerformanceSurface heat flux0.25-0.5 MW/m²
Neutron wall load0.5-2.0 MW/m²
Surface temperature limit< 50% of material melting point
Mechanical PerformanceThermal stress limit< 2/3 of allowable stress
Fatigue life> 10⁴ cycles
Pressure resistance> Coolant pressure × safety factor
Nuclear PropertiesNeutron transmissionAs high as possible
Impact on TBR< 0.05 reduction
Activation suppressionUse of low-activation materials
Plasma CompatibilitySputtering erosion< Allowable erosion/operating time
Impurity releasePrefer low-Z impurities
Tritium retentionBelow safety management limits

The total surface area of the first wall is approximately 680 m² for ITER and will exceed 1000 m² for future power reactors. Maintaining uniform cooling performance and mechanical integrity across this vast area is a major design challenge.

To understand the positional relationship of the first wall in a tokamak reactor, the following parameters are defined:

aFW=ap+δSOL+δgapa_{\text{FW}} = a_p + \delta_{\text{SOL}} + \delta_{\text{gap}}

Here, apa_p is the plasma minor radius, δSOL\delta_{\text{SOL}} is the scrape-off layer thickness (typically 3-10 cm), and δgap\delta_{\text{gap}} is the plasma-wall gap (5-10 cm). The distance between the first wall and the separatrix is determined by the balance between plasma control stability and plasma-wall interactions.

Because the first wall is separated from the outermost magnetic surface of the plasma, heat and particle loads are reduced compared to the divertor.

ParameterShielding BlanketTest Blanket
Surface heat flux (average)0.25 MW/m²0.25 MW/m²
Surface heat flux (maximum)0.5 MW/m²0.5 MW/m²
Neutron wall load (average)0.56 MW/m²0.78 MW/m²

Considering that the heat load of a household boiler is approximately 0.1 MW/m², it is clear that the first wall receives very large amounts of heat. Although smaller than the heat load to the divertor (over 10 MW/m²), the first wall accounts for approximately 80% of the plasma-facing area, so the total heat received cannot be ignored.

The surface heat flux qsurfq_{\text{surf}} to the first wall consists of multiple components:

qsurf=qrad+qCX+qconv+qγq_{\text{surf}} = q_{\text{rad}} + q_{\text{CX}} + q_{\text{conv}} + q_{\gamma}

The breakdown of each component is as follows.

The radiative heat flux qradq_{\text{rad}} from plasma is the sum of bremsstrahlung, line radiation, and synchrotron radiation:

qrad=Pbrem+Pline+PsyncAFWq_{\text{rad}} = \frac{P_{\text{brem}} + P_{\text{line}} + P_{\text{sync}}}{A_{\text{FW}}}

The bremsstrahlung power density is:

Pbrem=5.35×1037ne2ZeffTe1/2[W/m3]P_{\text{brem}} = 5.35 \times 10^{-37} n_e^2 Z_{\text{eff}} T_e^{1/2} \quad [\text{W/m}^3]

Here, nen_e is electron density [m⁻³], ZeffZ_{\text{eff}} is effective charge number, and TeT_e is electron temperature [keV]. Under typical ITER conditions (ne=1020n_e = 10^{20} m⁻³, Te=10T_e = 10 keV, Zeff=1.7Z_{\text{eff}} = 1.7), the bremsstrahlung power density is approximately 0.03 MW/m³.

Line radiation power from impurities is:

Pline=nenZLZ(Te)P_{\text{line}} = n_e n_Z L_Z(T_e)

LZ(Te)L_Z(T_e) is the radiative cooling coefficient, which depends on impurity species and temperature. For tungsten impurities, there is a strong radiation peak with LW1031L_W \sim 10^{-31} W·m³ in the electron temperature range of 1-5 keV.

Synchrotron radiation occurs when electrons move in the toroidal magnetic field:

Psync=6.21×1017neBT2TeTemec2[W/m3]P_{\text{sync}} = 6.21 \times 10^{-17} n_e B_T^2 T_e \frac{T_e}{m_e c^2} \quad [\text{W/m}^3]

Under high-temperature, high-field conditions this component cannot be ignored, but effective losses are reduced when wall reflectivity is high.

Charge Exchange Neutral Particle Heat Flux

Section titled “Charge Exchange Neutral Particle Heat Flux”

The heat flux qCXq_{\text{CX}} transported to the first wall by neutral particles produced in charge exchange reactions is:

qCX=ΓCXECXq_{\text{CX}} = \Gamma_{\text{CX}} \cdot \langle E_{\text{CX}} \rangle

The charge exchange flux ΓCX\Gamma_{\text{CX}} is:

ΓCX=nin0σvCXλCX\Gamma_{\text{CX}} = n_i n_0 \langle \sigma v \rangle_{\text{CX}} \cdot \lambda_{\text{CX}}

Here, n0n_0 is neutral particle density, σvCX\langle \sigma v \rangle_{\text{CX}} is the charge exchange reaction rate coefficient, and λCX\lambda_{\text{CX}} is the mean free path of neutral particles. Typical charge exchange neutral particle energies range from 0.1-10 keV.

Volumetric heating q˙vol\dot{q}_{\text{vol}} due to neutron irradiation decreases exponentially in the thickness direction of the first wall:

q˙vol(x)=q˙0exp(xλn)\dot{q}_{\text{vol}}(x) = \dot{q}_0 \exp\left(-\frac{x}{\lambda_n}\right)

Here, q˙0\dot{q}_0 is the volumetric heating rate at the surface, xx is the distance from the surface, and λn\lambda_n is the neutron attenuation length. The attenuation length for 14.1 MeV neutrons varies by material:

MaterialAttenuation length λn\lambda_n [cm]Surface heating rate (at 1 MW/m²) [MW/m³]
Beryllium1215
Steel825
Tungsten360

The total volumetric heating of the first wall is:

Qvol=0tFWq˙vol(x)dx=q˙0λn[1exp(tFWλn)]Q_{\text{vol}} = \int_0^{t_{\text{FW}}} \dot{q}_{\text{vol}}(x) \, dx = \dot{q}_0 \lambda_n \left[1 - \exp\left(-\frac{t_{\text{FW}}}{\lambda_n}\right)\right]

Here, tFWt_{\text{FW}} is the first wall thickness.

The temperature distribution in the first wall under steady-state conditions is derived from the one-dimensional heat conduction equation. With surface heat flux qsq_s and volumetric heating q˙\dot{q}:

kd2Tdx2=q˙(x)-k \frac{d^2 T}{dx^2} = \dot{q}(x)

Given boundary conditions of coolant-side temperature TcT_c and surface heat flux qsq_s, the surface temperature TsT_s is:

Ts=Tc+qstk+q˙0λn2k[1exp(tλn)tλnexp(tλn)]T_s = T_c + \frac{q_s \cdot t}{k} + \frac{\dot{q}_0 \lambda_n^2}{k} \left[1 - \exp\left(-\frac{t}{\lambda_n}\right) - \frac{t}{\lambda_n}\exp\left(-\frac{t}{\lambda_n}\right)\right]

In the typical temperature distribution for ITER’s first wall (beryllium surface, copper alloy structure), surface temperature is controlled within the range of 200-400°C.

Thermal stress due to temperature gradients directly affects the mechanical integrity of the first wall. The maximum thermal stress σth\sigma_{\text{th}} in a flat plate approximation is:

σth=EαΔT2(1ν)\sigma_{\text{th}} = \frac{E \alpha \Delta T}{2(1-\nu)}

Here, EE is Young’s modulus, α\alpha is the coefficient of linear thermal expansion, ν\nu is Poisson’s ratio, and ΔT\Delta T is the temperature difference.

The temperature difference due to surface heat flux qq is:

ΔT=qtk\Delta T = \frac{q \cdot t}{k}

Therefore, to suppress thermal stress, the following are desirable:

  • Thin first wall (small tt)
  • High thermal conductivity material (large kk)
  • Low thermal expansion material (small α\alpha)
  • Low Young’s modulus material (small EE)

The thermal stress parameter MM combining these parameters is used as an indicator for material selection:

M=kσy(1ν)EαM = \frac{k \sigma_y (1-\nu)}{E \alpha}

Here, σy\sigma_y is yield stress. Higher values of MM indicate tolerance to higher heat loads.

Charged particles in plasma normally move spiraling around magnetic field lines, but reach the first wall through the following processes:

  1. Charge exchange reactions: Plasma ions collide with neutral particles and lose their charge, and the resulting neutral particles, no longer confined by magnetic fields, strike the first wall
  2. Orbit loss of high-energy particles: High-energy ions such as alpha particles have large orbits and may directly collide with the first wall
  3. Diffusive transport: Part of the edge plasma diffuses across magnetic field lines and reaches the first wall

The energy of neutral particles produced by charge exchange reaches several hundred eV to several keV, causing sputtering when they collide with the first wall.

14.1 MeV DT neutrons are the first high-energy particles to reach the first wall. The neutron wall load Γn\Gamma_n is:

Γn=Pfusion×0.8AFW\Gamma_n = \frac{P_{\text{fusion}} \times 0.8}{A_{\text{FW}}}

Here, PfusionP_{\text{fusion}} is fusion power, 0.8 is the fraction of energy carried by neutrons (14.1 MeV / 17.6 MeV), and AFWA_{\text{FW}} is the first wall area.

Neutron fluence Φ\Phi is defined as cumulative irradiation:

Φ=Γntopfavail\Phi = \Gamma_n \cdot t_{\text{op}} \cdot f_{\text{avail}}

Here, topt_{\text{op}} is operating time and favailf_{\text{avail}} is availability. Neutron fluence of 2-3 MW·year/m² per year is anticipated for power reactors.

Surface materials for the first wall require strong compatibility with plasma. To minimize impact when entering the plasma, low atomic number (low-Z) materials are primarily selected.

Adopted as the first wall surface material for ITER.

ParameterValue
Atomic number4
Melting point1287 °C
Thermal conductivity200 W/(m·K)
Density1.85 g/cm³
Specific heat capacity1825 J/(kg·K)
Coefficient of linear thermal expansion11.3 × 10⁻⁶ /K
Young’s modulus287 GPa
Poisson’s ratio0.032

Advantages:

  • Low atomic number: Small radiation loss when entering plasma
  • Oxygen getter effect: Has property of reducing oxygen impurities
  • Relatively low tritium retention
  • High thermal conductivity and low density

Challenges:

  • Low melting point: Not suitable for high heat load regions
  • Toxicity: Handling requires caution
  • Steam reaction: Reacts with steam at high temperatures to generate hydrogen
Be+H2OBeO+H2\text{Be} + \text{H}_2\text{O} \rightarrow \text{BeO} + \text{H}_2

This reaction becomes significant above approximately 600°C, and the reaction rate is:

RBe-H2O=Aexp(EaRT)pH2OR_{\text{Be-H}_2\text{O}} = A \exp\left(-\frac{E_a}{RT}\right) \cdot p_{\text{H}_2\text{O}}

Here, A=1.4×108A = 1.4 \times 10^8 kg/(m²·s·Pa), Ea=108E_a = 108 kJ/mol, and pH2Op_{\text{H}_2\text{O}} is steam partial pressure.

Under neutron irradiation, beryllium generates helium through the following nuclear reactions:

9Be(n,2n)24He^9\text{Be}(n, 2n)2 \,^4\text{He} 9Be(n,α)6He6Li+β^9\text{Be}(n, \alpha)^6\text{He} \rightarrow ^6\text{Li} + \beta^-

The helium production rate is very high at approximately 500-1000 appm/dpa, causing swelling and embrittlement of the material. The swelling amount ΔV/V\Delta V / V shows temperature dependence:

ΔVV=kHecHef(T)\frac{\Delta V}{V} = k_{\text{He}} \cdot c_{\text{He}} \cdot f(T)

Here, kHek_{\text{He}} is the swelling coefficient per helium atom, cHec_{\text{He}} is helium concentration, and f(T)f(T) is a temperature-dependent function. Swelling is maximum in the 400-600°C range.

Tungsten may be used in high heat load regions.

ParameterValue
Atomic number74
Melting point3422 °C
Thermal conductivity173 W/(m·K)
Density19.3 g/cm³
Specific heat capacity132 J/(kg·K)
Coefficient of linear thermal expansion4.5 × 10⁻⁶ /K
Young’s modulus411 GPa
Sputtering threshold (D incidence)Approximately 300 eV
Recrystallization temperature1100-1400 °C

Tungsten has the highest melting point among metals and has low sputtering erosion, making it an excellent material, but due to its high atomic number, radiation loss is large when it enters the plasma.

The radiation loss power for tungsten concentration cWc_W in plasma is:

Prad,W=ne2cWLW(Te)P_{\text{rad},W} = n_e^2 c_W L_W(T_e)

The radiative cooling coefficient LWL_W of tungsten strongly depends on electron temperature, showing a peak of 103110^{-31} W·m³ in the 1-5 keV range. To maintain burning plasma, tungsten concentration must be kept below cW<105c_W < 10^{-5}.

Tungsten has a high ductile-brittle transition temperature (DBTT) (approximately 200-400°C for unirradiated material), and irradiation further increases the DBTT:

ΔTDBTT=kirrϕ1/2\Delta T_{\text{DBTT}} = k_{\text{irr}} \cdot \phi^{1/2}

Here, kirrk_{\text{irr}} is the irradiation hardening coefficient and ϕ\phi is neutron fluence. When operating temperature falls below DBTT, the risk of brittle fracture increases.

Detailed Comparison of Beryllium and Tungsten

Section titled “Detailed Comparison of Beryllium and Tungsten”
PropertyBerylliumTungstenNotes
Plasma radiation lossLowHighBe advantage
Melting point1287°C3422°CW advantage
Sputtering rateHighLowW advantage
Oxygen getterYesNoBe advantage
Tritium retentionLowLowEqual
Neutron multiplicationYesNoBe advantage (TBR)
ToxicityHighLowW advantage
MachinabilityDifficultDifficultEqual
CostHighMediumW advantage

The following materials are used for first wall structural materials:

  • Chromium zirconium copper (CuCrZr): Excellent thermal conductivity and durability (ITER)
  • Reduced activation ferritic steel (F82H, EUROFER97): Under development as structural material for prototype reactors
  • Stainless steel (SS316LN): Proven with existing technology

Reduced Activation Ferritic/Martensitic Steel (RAFM Steel)

Section titled “Reduced Activation Ferritic/Martensitic Steel (RAFM Steel)”

Properties of reduced activation ferritic/martensitic steel (RAFM steel) being developed as first wall structural material for power reactors:

ParameterF82HEUROFER97
CompositionFe-8Cr-2W-V-TaFe-9Cr-1W-V-Ta
Thermal conductivity28 W/(m·K)30 W/(m·K)
Coefficient of linear thermal expansion11.3 × 10⁻⁶ /K11.5 × 10⁻⁶ /K
Yield strength (RT)530 MPa550 MPa
Operating temperature range350-550°C350-550°C

These materials avoid elements that generate long-lived radioactive nuclides such as Mo and Nb, substituting them with low-activation elements such as W and Ta.

Water cooling is primarily adopted for first wall cooling.

The following cooling conditions are set for ITER:

  • Coolant: Water (3 MPa) or supercritical water (25 MPa)
  • Inlet temperature: 100 °C (water) / 280 °C (supercritical water)
  • Outlet temperature: 325 °C (water) / 510 °C (supercritical water)

Heat removal QQ is expressed by the following equation:

Q=ScvρcpΔT=m˙cpΔTQ = S_c \cdot v \cdot \rho \cdot c_p \cdot \Delta T = \dot{m} \cdot c_p \cdot \Delta T

Here, ScS_c is flow channel cross-sectional area, vv is flow velocity, ρ\rho is density, cpc_p is specific heat at constant pressure, ΔT\Delta T is temperature difference between inlet and outlet, and m˙\dot{m} is mass flow rate.

The heat transfer coefficient hh between coolant and wall surface depends on flow conditions. For forced convection, it is calculated from the Nusselt number NuNu:

h=NukfDhh = \frac{Nu \cdot k_f}{D_h}

Here, kfk_f is the thermal conductivity of the coolant and DhD_h is the hydraulic equivalent diameter.

Under turbulent conditions (Re>104Re > 10^4), the Dittus-Boelter correlation:

Nu=0.023Re0.8Pr0.4Nu = 0.023 \cdot Re^{0.8} \cdot Pr^{0.4}

Reynolds number ReRe and Prandtl number PrPr are:

Re=ρvDhμ,Pr=μcpkfRe = \frac{\rho v D_h}{\mu}, \quad Pr = \frac{\mu c_p}{k_f}

ITER’s first wall cooling system operates under conditions of Re105Re \sim 10^5, h104h \sim 10^4 W/(m²·K).

Advantages of water cooling include high heat transfer coefficient and proven existing technology. However, the following challenges exist:

  • Tritium permeation: Tritium permeation management to water is necessary
  • Corrosion: Corrosion due to radiolysis of water in radiation environment
  • Pressure constraints: Mechanical constraints from high-pressure operation

Radiolysis of cooling water produces the following reactions:

H2Oγ,nH+OHH2+H2O2+O2\text{H}_2\text{O} \xrightarrow{\gamma, n} \text{H}^{\bullet} + \text{OH}^{\bullet} \rightarrow \text{H}_2 + \text{H}_2\text{O}_2 + \text{O}_2

The generated hydrogen peroxide H2O2\text{H}_2\text{O}_2 and dissolved oxygen promote corrosion of piping materials. Water chemistry management is required to control dissolved oxygen concentration at ppb levels.

For power reactors, helium gas cooling is being considered for improved thermal efficiency through high-temperature operation:

ParameterWater cooling (ITER)Helium cooling (future reactors)
Coolant pressure3-4 MPa8-10 MPa
Inlet temperature100°C300°C
Outlet temperature150°C500°C
Thermal efficiency30-33%40-45%

The heat transfer coefficient of helium cooling is about 1/10 that of water, so extended heat transfer surfaces and turbulence-promoting structures are required. The following design methods are used for heat transfer enhancement:

henhanced=hsmoothηfinAfinAbaseh_{\text{enhanced}} = h_{\text{smooth}} \cdot \eta_{\text{fin}} \cdot \frac{A_{\text{fin}}}{A_{\text{base}}}

Here, ηfin\eta_{\text{fin}} is fin efficiency and Afin/AbaseA_{\text{fin}}/A_{\text{base}} is surface area expansion ratio.

The following types of flow channel cross-sections are available for the first wall:

  1. Tube type: Structure with many circular tubes joined together. Circular cross-section is ideal for coolant internal pressure
  2. Rib type: Integrated structure with rectangular cross-section channels. Can reduce coolant equivalent thickness, improving tritium breeding performance
  3. Corrugated type: Structure with corrugated plates joined to flat plates. Can reduce coolant equivalent thickness but has manufacturability challenges
ItemIntegrated typeSeparate type
StructureFirst wall and blanket are integratedCan be exchanged independently
TBRThin plate thickness with small impactThick plate thickness with large impact
MaintenanceSimultaneous exchange requiredCan be exchanged separately

The integrated type is currently mainstream from the perspectives of structural simplicity and tritium breeding ratio.

When plasma particles collide with the first wall, surface atoms are ejected through momentum transfer.

This depends on incident ion energy and material atomic number. The sputtering yield (number of atoms released per incident ion) has an energy threshold, and for high-Z materials like tungsten, the threshold is high so sputtering does not occur with low-energy ions.

Physical sputtering yield YY is described by a semi-empirical model:

Y(E)=Qsn(ε)[1(EthE)2/3](1EthE)2Y(E) = Q \cdot s_n(\varepsilon) \cdot \left[1 - \left(\frac{E_{\text{th}}}{E}\right)^{2/3}\right] \cdot \left(1 - \frac{E_{\text{th}}}{E}\right)^2

Here, QQ is a material constant, sn(ε)s_n(\varepsilon) is reduced nuclear stopping power, EthE_{\text{th}} is threshold energy, and EE is incident energy.

The threshold energy is approximated by:

Eth=Esg(1g)E_{\text{th}} = \frac{E_s}{g(1-g)}

Here, EsE_s is surface binding energy and g=4M1M2/(M1+M2)2g = 4M_1 M_2 / (M_1 + M_2)^2 is the kinematic factor (M1M_1: incident particle mass, M2M_2: target atom mass).

For beryllium, the sputtering threshold is approximately 10 eV, which is low, but for tungsten it increases to approximately 300 eV.

Incident particleBeryllium threshold [eV]Tungsten threshold [eV]
H12450
D9220
T8150
He20120

This is a phenomenon specific to carbon materials, where hydrogen isotope ions react chemically to produce hydrocarbons such as methane (CH₄) and ethylene (C₂H₄), causing material erosion. The sputtering rate is maximum around 500 °C, and erosion can be about 10 times that of physical sputtering.

Temperature dependence of chemical sputtering rate:

Ychem(T)=Y0exp[EakB(1T1Tmax)2/σ2]Y_{\text{chem}}(T) = Y_0 \cdot \exp\left[-\frac{E_a}{k_B}\left(\frac{1}{T} - \frac{1}{T_{\text{max}}}\right)^2 / \sigma^2\right]

Hydrocarbons generated by this reaction dissociate in the plasma and form co-deposited layers containing tritium. Tritium accumulation in these layers poses a major safety management challenge, so carbon material use is restricted in ITER.

Annual erosion amount Δt\Delta t is:

Δt=YΓMρNAtyear\Delta t = \frac{Y \cdot \Gamma \cdot M}{\rho \cdot N_A} \cdot t_{\text{year}}

Here, Γ\Gamma is particle flux [m⁻²s⁻¹], MM is atomic mass, ρ\rho is density, NAN_A is Avogadro’s number, and tyeart_{\text{year}} is annual operating seconds.

Under ITER conditions, annual sputtering erosion of the beryllium first wall is estimated at approximately 0.1-1 mm/year.

Atoms released by sputtering are ionized in the plasma, transported along magnetic field lines, and then redeposited elsewhere. This redeposition process significantly affects the surface condition and lifetime of the first wall.

The growth rate of redeposited layers RdepR_{\text{dep}} is:

Rdep=ΓdepMρNAR_{\text{dep}} = \frac{\Gamma_{\text{dep}} \cdot M}{\rho \cdot N_A}

Here, Γdep\Gamma_{\text{dep}} is deposited particle flux. Redeposited layers often have porous structures and show different properties from bulk materials:

PropertyBulk materialRedeposited layer
Density100% of theoretical density50-80%
Thermal conductivityStandard value1/5-1/10
Mechanical strengthStandard valueSignificantly reduced
Tritium retentionLowHigh

The reduced thermal conductivity of redeposited layers creates local overheating risks and causes layer delamination and particle generation.

Spatial Distribution of Erosion/Redeposition

Section titled “Spatial Distribution of Erosion/Redeposition”

The first wall surface is divided into erosion-dominated and redeposition-dominated regions depending on local plasma conditions. Net erosion rate EnetE_{\text{net}} is:

Enet=EgrossRdep=YΓionSdepΓneutralE_{\text{net}} = E_{\text{gross}} - R_{\text{dep}} = Y \cdot \Gamma_{\text{ion}} - S_{\text{dep}} \cdot \Gamma_{\text{neutral}}

In tokamaks, generally:

  • Near the outer midplane: Erosion-dominated
  • Upper and lower poloidal regions: Redeposition-dominated
  • Near divertor: Mixed region

In ITER, which uses both beryllium and tungsten, mixed layers form on surfaces through sputtering and redeposition. In Be-W mixed layers:

BexW1x(x=0.20.8)\text{Be}_x\text{W}_{1-x} \quad (x = 0.2 \sim 0.8)

Intermetallic compounds (Be₂W, Be₁₂W, etc.) form in this composition range, reducing the melting point below that of bulk materials. The melting point of Be₂W is approximately 2200°C, significantly lower than tungsten (3422°C). This phenomenon increases the risk of local melting under high heat loads.

Plasma irradiation causes characteristic morphological changes on first wall surfaces:

  1. Nanostructure formation: “Fuzz” structures formed on tungsten surfaces

    • Generated by helium ion irradiation
    • Highly porous nanofibrous structure
    • Significant reduction in thermal conductivity
  2. Surface roughening: Irregularity formation from ion irradiation

    • Due to angular dependence of sputtering yield
    • Increased gas retention due to surface area increase
  3. Bubbles and blisters: Surface bulging from gas accumulation

Tungsten fuzz formation conditions:

  • Helium ion energy > 20 eV
  • Surface temperature 1000-2000 K
  • Helium fluence > 10²⁴ m⁻²

Incident gas atoms precipitate as bubbles near the material surface and grow with irradiation dose. Eventually, surface blisters rupture, causing surface delamination and flaking.

Erosion thickness ϵ\epsilon due to blistering is:

ϵ=Rϕcϕ\epsilon = \frac{R}{\phi_c} \cdot \phi

Here, RR is blister skin thickness, ϕc\phi_c is critical irradiation dose, and ϕ\phi is irradiation dose. The critical dose decreases as material temperature increases.

The critical condition for blister formation is from the balance between bubble internal pressure and surface tension:

pbubble=2γr+σyp_{\text{bubble}} = \frac{2 \gamma}{r} + \sigma_y

Here, γ\gamma is surface energy, rr is bubble radius, and σy\sigma_y is yield stress.

During plasma disruptions, large amounts of energy are deposited on the first wall in short times of 1-10 ms during the thermal quench phase. Heat loads reach 1-10 MJ/m², causing surface melting and evaporation.

Erosion due to evaporation is:

Δm=QdepQmeltLv\Delta m = \frac{Q_{\text{dep}} - Q_{\text{melt}}}{L_v}

Here, QdepQ_{\text{dep}} is deposited energy density, QmeltQ_{\text{melt}} is energy required for melting, and LvL_v is latent heat of vaporization.

For beryllium, evaporation erosion of approximately 10 μm is predicted for a heat load of 1 MJ/m². Considering the number of disruptions expected during ITER’s operating period (several hundred), cumulative erosion may reach several mm.

Neutron irradiation displaces material atoms from lattice positions. Damage is expressed in dpa (displacements per atom):

dpa=0tEminEmaxϕ(E)σd(E)ν(E)dEdt\text{dpa} = \int_0^{t} \int_{E_{\min}}^{E_{\max}} \phi(E) \cdot \sigma_d(E) \cdot \nu(E) \, dE \, dt

Here, ϕ(E)\phi(E) is neutron flux, σd(E)\sigma_d(E) is displacement cross-section, and ν(E)\nu(E) is cascade multiplication factor.

The dpa accumulation rate in the first wall is:

Materialdpa rate (at 1 MW/m²) [dpa/year]
Beryllium10
Steel8
Tungsten5

For power reactors, 20-30 dpa/year and 100-300 dpa over lifetime (5-10 years) of damage is predicted.

Neutron irradiation generates the following defects:

  1. Frenkel pairs: Pairs of interstitial atoms and vacancies
  2. Dislocation loops: Structures from defect agglomeration
  3. Voids/bubbles: Vacancy agglomeration or gas atom precipitation
  4. Precipitates: Agglomeration of transmutation products

The time evolution of defect concentration is described by rate equations:

dcvdt=GvRivcicvkv2Dvcv\frac{dc_v}{dt} = G_v - R_{iv} c_i c_v - k_v^2 D_v c_v dcidt=GiRivcicvki2Dici\frac{dc_i}{dt} = G_i - R_{iv} c_i c_v - k_i^2 D_i c_i

Here, cvc_v, cic_i are vacancy and interstitial concentrations, GG is generation rate, RivR_{iv} is recombination coefficient, k2k^2 is sink strength, and DD is diffusion coefficient.

Nuclear reactions between fusion neutrons and materials generate helium and hydrogen. Helium generation rate by (n, α) reactions:

MaterialHe generation rate (at 1 MW/m²) [appm/dpa]
Beryllium500-1000
Steel10-15
Tungsten0.5-1

Main reactions for helium generation in steel:

56Fe(n,α)53Cr^{56}\text{Fe}(n, \alpha)^{53}\text{Cr} 58Ni(n,α)55Fe^{58}\text{Ni}(n, \alpha)^{55}\text{Fe}

Helium accumulates in materials and segregates to grain boundaries, causing high-temperature grain boundary embrittlement. When critical helium concentration is exceeded, grain boundary strength decreases rapidly:

σGB=σGB,0exp(cHecHe)\sigma_{\text{GB}} = \sigma_{\text{GB},0} \exp\left(-\frac{c_{\text{He}}}{c_{\text{He}}^*}\right)

Irradiation defects act as obstacles to dislocation motion, causing material hardening. The increase in yield stress Δσy\Delta \sigma_y is:

Δσy=MμbjNjdj\Delta \sigma_y = M \mu b \sqrt{\sum_j N_j d_j}

Here, MM is Taylor factor, μ\mu is shear modulus, bb is Burgers vector, and NjN_j, djd_j are number density and diameter of various defects.

Irradiation embrittlement increases the ductile-brittle transition temperature (DBTT):

ΔTDBTTΔσy\Delta T_{\text{DBTT}} \propto \Delta \sigma_y

For reduced activation ferritic steel, a DBTT increase of 100-200°C has been reported at 10 dpa irradiation.

Void formation from vacancy agglomeration causes volumetric expansion of materials. Swelling amount is:

ΔVV=4π3iNiri3\frac{\Delta V}{V} = \frac{4\pi}{3} \sum_i N_i r_i^3

Austenitic stainless steel shows high swelling rates (several %/dpa), making its use difficult in fusion reactors after ITER. Ferritic/martensitic steel has excellent swelling resistance (< 0.1 %/dpa) and is expected as structural material for power reactors.

Tritium (T) retention in first wall materials occurs through the following processes:

  1. Implantation retention: Direct implantation by ion incidence
  2. Diffusion penetration: Absorption and diffusion from gas phase
  3. Defect trapping: Trapping at irradiation defects
  4. Co-deposition: Incorporation into sputtering redeposition layers

Total tritium retention ITI_T is:

IT=Iimplant+Idiffusion+Itrap+IcodepositI_T = I_{\text{implant}} + I_{\text{diffusion}} + I_{\text{trap}} + I_{\text{codeposit}}

Tritium diffusion into materials follows Fick’s law:

Ct=D2CjkjC(NjCjT)+jpjCjT\frac{\partial C}{\partial t} = D \nabla^2 C - \sum_j k_j C (N_j - C_j^T) + \sum_j p_j C_j^T

Here, CC is dissolved concentration, DD is diffusion coefficient, kjk_j, pjp_j are trapping/detrapping rates for various traps, NjN_j is trap density, and CjTC_j^T is trapped concentration.

Temperature dependence of diffusion coefficient:

D=D0exp(EDkBT)D = D_0 \exp\left(-\frac{E_D}{k_B T}\right)

Solubility:

S=S0exp(ESkBT)S = S_0 \exp\left(-\frac{E_S}{k_B T}\right)

Diffusion and solubility parameters for representative materials:

MaterialD0D_0 [m²/s]EDE_D [eV]S0S_0 [mol/(m³·Pa^0.5)]ESE_S [eV]
Be3×10⁻⁹0.282×10⁻²0.9
W4×10⁻⁷0.3910⁻³1.0
Steel2×10⁻⁷0.140.30.1

Tritium permeation flux through the wall JTJ_T is:

JT=DStpTJ_T = \frac{D \cdot S}{t} \cdot \sqrt{p_T}

Here, tt is wall thickness and pTp_T is tritium partial pressure. It is characterized by permeability Φ=DS\Phi = D \cdot S.

Oxide coatings (Al₂O₃, Er₂O₃, etc.) are being considered for permeation prevention. The permeation reduction factor (PRF) by coating is:

PRF=JbareJcoated=1+ΦbasetcoatΦcoattbase\text{PRF} = \frac{J_{\text{bare}}}{J_{\text{coated}}} = 1 + \frac{\Phi_{\text{base}} \cdot t_{\text{coat}}}{\Phi_{\text{coat}} \cdot t_{\text{base}}}

Under ideal conditions, PRF > 1000 is achieved, but maintaining performance in real environments (irradiation, thermal cycling) is a challenge.

A safety management limit of 700 g is set for tritium inventory in ITER. To meet this limit:

  • Adoption of low tritium retention material (beryllium)
  • Restrictions on carbon material use
  • Periodic tritium removal operations
  • Online monitoring

are implemented.

ITER’s first wall is designed as part of the shielding blanket. Main specifications are as follows:

ParameterValue
Total surface area680 m²
Number of blanket modules440
Number of first wall panelsApproximately 1800
Surface materialBeryllium
Structural materialCuCrZr / 316L(N)-IG
CoolantWater (4 MPa, 70°C→120°C)

ITER’s first wall panel has the following layered structure:

  1. Beryllium armor (8-10 mm)
  2. CuCrZr heat sink (10 mm)
  3. 316L(N) stainless steel structure (20 mm)

HIP (Hot Isostatic Pressing) bonding is used for joining beryllium and copper alloy. The integrity of the bonding interface must be maintained against thermal cycling and neutron irradiation during operation.

Residual stress σres\sigma_{\text{res}} at the bonding interface from thermal expansion coefficient difference:

σres=Eeff(αBeαCu)ΔT1+EBetBe/ECutCu\sigma_{\text{res}} = \frac{E_{\text{eff}} (\alpha_{\text{Be}} - \alpha_{\text{Cu}}) \Delta T}{1 + E_{\text{Be}} t_{\text{Be}} / E_{\text{Cu}} t_{\text{Cu}}}

Two types of panels are used for ITER first wall depending on heat load conditions:

TypeSurface heat fluxBeryllium thicknessInstallation location
Standard (NHF)2 MW/m²8 mmMost areas
Enhanced (EHF)4.7 MW/m²10 mmHigh heat load regions

Enhanced panels employ a finger structure to improve heat transfer performance.

The first wall cooling system operates under the following design conditions:

  • Coolant: Deionized water
  • Pressure: 4 MPa
  • Inlet temperature: 70°C
  • Outlet temperature: 120°C
  • Total flow rate: Approximately 1000 kg/s

Independent cooling loops are configured for each blanket module, minimizing impact on other modules in case of cooling loss in one module.

ITER’s first wall/blanket is exchanged using a remote handling (RH) system. Since the interior of the fusion reactor is strongly activated after DT operation, direct human access is impossible, and all maintenance work is performed by remote operation.

ITER’s in-vessel remote maintenance system consists of the following elements:

  1. Blanket Remote Handling System (BRHS)

    • Multi-jointed manipulator inserted through vacuum vessel ports
    • Maximum payload: 5 tons
    • Positioning accuracy: ±2 mm
    • Working range: Entire vacuum vessel interior
  2. Divertor Remote Handling System (DRHS)

    • Access from lower ports
    • Designed exclusively for cassette exchange
  3. In-vessel inspection system

    • Cameras and lighting for visual inspection
    • Surface erosion evaluation by laser measurement
    • Thermal integrity confirmation by infrared camera
  1. Vacuum vessel evacuation and purge (1-2 days)
  2. Port plug removal (1 day per port)
  3. Cooling pipe cutting and capping
  4. Module fixing bolt removal
  5. Module gripping and removal by manipulator
  6. New module insertion and positioning
  7. Fixing bolt fastening
  8. Cooling pipe reconnection and leak testing
  9. Port plug restoration

Main parameters of blanket modules:

  • Total number: 440 modules
  • Weight: Approximately 4-4.5 tons/module
  • Exchange time: Approximately 2-3 weeks/module
  • Full exchange: Approximately 6-12 months

Remote maintenance systems must operate in the following radiation environment:

ParameterImmediately after shutdownAfter 2 weeksAfter 1 year
γ dose rate10⁴ Gy/h10³ Gy/h10² Gy/h
Contact dose rate10⁵ mSv/h10⁴ mSv/h10³ mSv/h

In this environment:

  • Radiation degradation countermeasures for electronics (radiation-resistant design)
  • Prevention of optical system browning (special glass, periodic replacement)
  • Countermeasures for lubricant radiolysis (solid lubrication, special grease)
  • Cable insulation material degradation management

are necessary. Remote maintenance equipment is designed for a cumulative dose target of 10⁶ Gy (lifetime).

First wall replacement timing is determined from the following indicators:

  1. Surface erosion amount
tremain=tinitial0tt˙erosion(t)dt>tmint_{\text{remain}} = t_{\text{initial}} - \int_0^{t} \dot{t}_{\text{erosion}}(t') \, dt' > t_{\text{min}}
  1. Neutron damage amount
dpaaccumulated<dpalimit\text{dpa}_{\text{accumulated}} < \text{dpa}_{\text{limit}}
  1. Tritium retention amount
IT<IT,limitI_T < I_{T,\text{limit}}
  1. Thermal fatigue damage
Dfatigue=jnjNf,j<1(Miner’s rule)D_{\text{fatigue}} = \sum_j \frac{n_j}{N_{f,j}} < 1 \quad (\text{Miner's rule})

One to two full replacements are planned during ITER’s DT operation period.

For power reactors (DEMO and beyond), the following points become more demanding compared to ITER:

ParameterITERDEMOCommercial reactor
Neutron wall load0.56 MW/m²2-3 MW/m²3-5 MW/m²
Annual availability10%50%80%
Lifetime dpa350-100150-200
Operating time1000 secondsSteady-stateSteady-state

To address the high neutron wall load (2-5 MW/m²) of power reactors:

  1. Development of high thermal conductivity materials
  2. Advanced cooling structures (microchannels, porous bodies)
  3. Application of composite materials
  4. Functionally graded materials

are being considered. The allowable range when surface heat flux increases:

qmax=k(TmeltTcoolant)tarmorηsafetyq_{\text{max}} = \frac{k \cdot (T_{\text{melt}} - T_{\text{coolant}})}{t_{\text{armor}}} \cdot \eta_{\text{safety}}

Expansion of Irradiation Materials Database

Section titled “Expansion of Irradiation Materials Database”

Current material irradiation data extends only to about 100 dpa at most, and there is uncertainty in predicting behavior under power reactor conditions (150-200 dpa). In particular:

  • Helium accumulation and grain boundary embrittlement
  • Effects of transmutation products
  • Combined damage (irradiation + thermal cycling)

require data expansion. Material irradiation testing at 14 MeV neutron source facilities (IFMIF-DONES, etc.) is planned.

The following advanced materials are under development for future reactors:

  1. ODS steel (Oxide Dispersion Strengthened steel)

    • Strengthening by nano-sized oxide particles (Y₂O₃-Ti-O)
    • Improved creep strength and irradiation resistance
    • Upper temperature limit extended to 650°C or higher
  2. SiC/SiC composite materials

    • Low activation characteristics
    • High-temperature strength
    • Operating temperatures above 1000°C
    • Challenges: Gas-tightness, joining technology
  3. Tungsten alloys

    • Recrystallization suppression (K-doped, La₂O₃ dispersed)
    • Toughness improvement (ReW alloy)
    • Irradiation embrittlement suppression

For power reactors, long lifetime of the first wall/blanket is required from an economic perspective:

COECFW/BLKtlifefavail\text{COE} \propto \frac{C_{\text{FW/BLK}}}{t_{\text{life}} \cdot f_{\text{avail}}}

Here, COE is cost of electricity, CFW/BLKC_{\text{FW/BLK}} is first wall/blanket cost, tlifet_{\text{life}} is lifetime, and favailf_{\text{avail}} is availability.

Target values:

  • Lifetime 5 years or more
  • Exchange time within 3 months
  • Availability 80% or more

are set. Achieving these requires both improved material durability and more efficient maintenance operations.

Various advanced first wall concepts are being researched for future fusion power reactors. These aim to develop the technology demonstrated in ITER toward higher performance and reliability.

A concept for replenishing plasma-facing surface erosion during operation.

A method using liquid metals such as lithium (Li) or tin (Sn) flowing over the first wall surface as the plasma-facing surface instead of solid materials:

Advantages:

  • Self-replenishment of sputtering erosion
  • Response to high heat loads (evaporative cooling effect)
  • Tritium capture (for Li)
  • Self-recovery from radiation damage

Challenges:

  • Impurity contamination into plasma
  • MHD effects in magnetic fields
  • Liquid metal circulation and recovery systems
  • Material compatibility at high temperatures

The plasma impurity effect of liquid lithium is evaluated by evaporation flux Γevap\Gamma_{\text{evap}}:

Γevap=pLi(T)2πmLikBT\Gamma_{\text{evap}} = \frac{p_{\text{Li}}(T)}{\sqrt{2\pi m_{\text{Li}} k_B T}}

At surface temperatures below 450°C, impurity contamination from evaporation can be kept within acceptable limits.

A method of supplying material lost in erosion-dominated regions from elsewhere. Periodic surface restoration by plasma spraying of tungsten powder is being considered.

Materials with continuously varying composition or structure from surface to substrate. Relaxes interface stress from thermal expansion coefficient mismatch and improves joint reliability.

By continuous composition change from tungsten (surface) to copper (substrate):

αFGM(x)=αW+(αCuαW)f(x)\alpha_{\text{FGM}}(x) = \alpha_W + (\alpha_{Cu} - \alpha_W) \cdot f(x)

Mitigates sharp thermal stress concentration at interfaces. Manufacturing methods include powder metallurgy, thermal spraying, and PVD/CVD methods.

Combination of tungsten and reduced activation ferritic steel ensures structural material compatibility for power reactors.

A concept for improving material properties through nanoscale microstructures.

By refining grain size to below 100 nm:

  • Lowering of ductile-brittle transition temperature
  • Increase of recrystallization temperature
  • Promotion of irradiation defect annihilation at grain boundaries

are expected. From the Hall-Petch equation:

σy=σ0+kyd1/2\sigma_y = \sigma_0 + k_y d^{-1/2}

strength also improves with grain size dd refinement.

Materials with nano-sized oxide particles (Y₂O₃, La₂O₃, etc.) dispersed in a tungsten matrix. Dispersed particles function as sinks for irradiation defects, suppressing irradiation embrittlement.

A concept where materials themselves repair irradiation damage.

Ceramics with layered structures such as Ti₃SiC₂ and Ti₂AlC, showing special defect recovery mechanisms:

  • Stress relaxation through easy interlayer slip
  • Promotion of defect annihilation at high temperatures
  • Metallic thermal and electrical conductivity

Alloys mixing five or more elements in equimolar ratios, showing properties not found in conventional materials:

  • Improved radiation damage resistance (lattice strain effect)
  • Maintenance of high-temperature strength
  • Excellent corrosion resistance

WTaCrVTi-based high entropy alloys are being researched for fusion applications.

High-performance cooling structures to replace conventional tubular cooling channels.

Arranges many microchannels on the order of several hundred μm to expand heat transfer area. Heat transfer coefficient improvement:

hmicro=hmacro(DmacroDmicro)0.2h_{\text{micro}} = h_{\text{macro}} \cdot \left(\frac{D_{\text{macro}}}{D_{\text{micro}}}\right)^{0.2}

achieves 2-5 times heat removal performance with the same flow rate.

Flows coolant through porous structures of metal foam or sintered bodies, achieving high-efficiency cooling utilizing internal surface area.

Incorporates high thermal conductivity heat pipe structures into the first wall to distribute localized high heat loads.

Design Optimization by Computational Materials Science

Section titled “Design Optimization by Computational Materials Science”

Efforts are advancing to optimize material design from the atomic level using first-principles calculations and molecular dynamics simulations.

Continuously models irradiation damage evolution at multiple scales:

  1. First-principles calculations (electronic states, interatomic potentials)
  2. Molecular dynamics (cascade damage, defect formation)
  3. Kinetic Monte Carlo (defect diffusion, aggregation)
  4. Rate theory models (microstructure evolution)
  5. Finite element method (mechanical property changes)

This hierarchical approach predicts material behavior in high-irradiation regimes where experiments are difficult.

Research combining large-scale materials databases with machine learning algorithms to efficiently search for optimal material compositions and structures is beginning.

The first wall is critical equipment directly related to fusion reactor safety, requiring strict regulations and safety evaluation.

Safety functions performed by the first wall:

  1. Confinement of radioactive materials

    • Tritium retention within vacuum vessel
    • Prevention of activated product dispersion
  2. Decay heat removal

    • Residual heat removal after shutdown
    • Coordination with passive safety systems
  3. Contribution to reactivity control

    • Natural shutdown through plasma impurities

Events to be considered in first wall design:

Event categorySpecific examplesExpected frequency
Normal operationSteady-state heat load, pulse operationContinuous
Anticipated operational occurrencesELMs, small-scale disruptions10³-10⁵ times/lifetime
Design basis accidentsLarge-scale disruptions10-10² times/lifetime
Design extension conditionsLoss of coolant, vacuum breach< 10⁻² /year

Temperature rise is evaluated when first wall cooling is lost. Temperature rise rate under adiabatic conditions:

dTdt=q˙decayρcp\frac{dT}{dt} = \frac{\dot{q}_{\text{decay}}}{\rho c_p}

Here, q˙decay\dot{q}_{\text{decay}} is decay heat generation rate. For ITER’s first wall, decay heat immediately after shutdown is approximately 1% of normal operation, and passive heat removal by natural convection and radiation maintains temperatures below material melting points.

Hydrogen generation from beryllium-steam reactions and radiolysis is important from an explosion risk perspective:

m˙H2=ABeRBe-H2O(T)xexposed\dot{m}_{H_2} = A_{\text{Be}} \cdot R_{\text{Be-H}_2\text{O}}(T) \cdot x_{\text{exposed}}

Here, ABeA_{\text{Be}} is beryllium exposed area and xexposedx_{\text{exposed}} is the fraction of reactable surface.

The following countermeasures are taken to prevent hydrogen concentration from reaching explosive limits (4-75 vol%):

  • Dilution through vacuum vessel internal volume
  • Inert gas purge system
  • Catalytic recombiners

Tritium retained in the first wall is evaluated as a release source term in accidents. Release amount is:

RT=ITfrelease(T,t)R_T = I_T \cdot f_{\text{release}}(T, t)

Here, freleasef_{\text{release}} is a temperature and time-dependent release fraction. ITER’s tritium inventory limit (700 g) is set based on this evaluation.

Activation of first wall materials affects waste classification at decommissioning. Achieving clearance levels after 100 years is an important design goal in material selection:

A(t)=A0ifiexp(ln2T1/2,it)A(t) = A_0 \sum_i f_i \exp\left(-\frac{\ln 2}{T_{1/2,i}} t\right)

Low activation materials (RAFM steel, SiC/SiC, V alloys) exclude elements that produce long-lived nuclides such as Mo, Nb, and Ni, aiming to achieve clearance levels in about 100 years.