Paper 11
Paper 12
Paper 1
Paper 5
Paper 9
100

Explain how solvent quality affects PEG conformation (coil vs. helix), and how this is measured experimentally.

Good Solvent (e.g., water, ethanol):

  • Strong PEG–solvent hydrogen bonding
  • Chains adopt expanded coil conformations to maximize entropy
  • Radius of gyration increases
  • Chain flexibility is maximized → isotropic scattering

Poor Solvent or Mixed Solvent (e.g., isobutyric acid, water/IBA):

  • Solvent–PEG interactions are weaker than PEG–PEG interactions
  • PEG chains collapse or self-interact
  • In certain conditions (e.g., specific PEG length in IBA/H₂O), this collapse leads to chiral helical ordering

SANS:

  • Provides real-space information about chain size, shape, and fractal structure
  • Guinier–Porod analysis distinguishes between:

    • Coils: q^−2 or q^−5/3 behavior
    • Rod-like helices: q^−1 or sharp transition to Porod regime
  • Radius of gyration (Rg) indicates coil expansion/collapse

Optical Rotation via Polarimetry:

  • Helical PEG displays optical activity, especially when doped with chiral additives (e.g., camphor sulfonic acid)
  • Optical rotation increases with:

    • PEG chain length
    • Degree of helical ordering
    • Chiral dopant concentration
100

How can sin²ψ analysis be used to quantify lattice strain in nanostructured materials? What are the limitations?

Using Bragg peak shifts at multiple ψ tilts, the sin²ψ method quantifies in-plane and out-of-plane strain components by fitting strain as a function of sin²ψ. It assumes a uniform strain state and enables extraction of the full strain tensor via tensor rotation.

Limitations include:
– Assumes homogeneous strain—not valid for nanostructures with gradients
– Peak broadening and asymmetry from inhomogeneous strain distort results
– Loses accuracy in ultra-thin films or near interfaces
– Cannot resolve strain distributions, only average values.

100

This parameter, denoted χ, quantifies the thermodynamic incompatibility between blocks in a copolymer.

Flory-Huggins interaction parameter

100

After heat treatment at 650°C, this microstructural feature became more homogeneous, reducing internal stress and contributing to increased ductility.

cellular substructure and residual stress field

100

This type of chain in the amorphous region—responsible for stress transfer between lamellae—must degrade for fragmentation into nanoplastics to occur under quiescent conditions.

tie-molecules (including bridges and bridging entanglements)

200

How does chirality in a polymer-solvent system influence optical rotation? Compare with Ising model predictions.

  • Optical rotation arises when a chiral medium rotates the plane of polarized light due to its asymmetric molecular organization.
  • In the case of PEG in water–IBA mixtures doped with a chiral molecule (e.g., camphorsulfonic acid):

    • PEG can adopt left- or right-handed helices, depending on the chiral field imposed by the solvent.
    • The balance between enthalpic interactions (e.g., hydrogen bonding) and entropic preferences (e.g., chain flexibility) determines the induced handedness.

Key Dependencies:

  • Polymer chain length: Longer chains more readily support helical domains.
  • Chiral dopant concentration: Higher levels promote stronger helicity bias.
  • Solvent quality: Marginal solvents (e.g., near a coil–helix transition) amplify chiral effects by destabilizing random coils.

The Ising model, classically used for spin systems, has been adapted to describe chiral symmetry breaking and helix formation in polymers. Each monomeric unit can be considered to exist in one of two chiral states (e.g., left-handed or right-handed helix), analogous to spin-up or spin-down states.

1D Helical Ising-like Model:

  • Each monomer state si=±1 (left or right helix)
  • Hamiltonian:

H=−J∑sisi+1−h∑si


    • J: cooperativity parameter (interaction between neighboring units)
    • h: external chiral field (e.g., due to chiral dopant)

Predictions:

  • In absence of chiral field (h=0), the polymer shows no net helicity (optical rotation ≈ 0) unless spontaneous symmetry breaking occurs.
  • With small chiral field (h≠0), optical rotation grows nonlinearly with h, especially if cooperativity (J) is high → longer helical domains.
  • Critical behavior: Near the transition, small changes in h cause large shifts in helicity and optical rotation (sharp onset).

Chirality in polymer–solvent systems induces asymmetric chain conformations that manifest in measurable optical rotation. This effect is cooperative and nonlinear, as captured by the Ising model. The model explains why short PEGs show weak rotation (low J, low N), while longer, cooperative chains in a biased field exhibit macroscopic chiral signatures — an essential mechanism in helical amplification and chiral sensing.

200

What is the significance of direction-dependent peak broadening in XRD of thin films? How do you decouple strain effects?

Direction-dependent peak broadening indicates strain anisotropy or crystallite size anisotropy. In XRD, if broadening varies with diffraction angle, it's often due to strain gradients or coherent domain deformation.

To decouple strain and size effects:
Use the Williamson–Hall method, where:
Δq_total = Δq_size + Δq_strain
Size broadening ∝ 1/cosθ, strain broadening ∝ tanθ.
Plotting βcosθ vs sinθ allows separation of slope (strain) and intercept (size).

200

When the block volume fractions are nearly equal and χN is high, this type of bicontinuous morphology can form

gyroid (Ia3d) structure

200

Unlike traditional wrought steel, AM austenitic stainless steel exhibits this feature due to rapid solidification, which is partially preserved after sub-recrystallization heat treatments.

fine cellular-dendritic microstructure aligned along the build direction

200

This specific measurable event in tensile testing correlates with the onset of nanoplastic release during PET degradation and matches the induction period observed in light scattering data.

drop in maximum stress/failure strength

300

Discuss the entropic penalties of chain confinement in a poor solvent. Relate to polymer conformation transitions

  • In good solvents, polymer segments experience favorable interactions with the solvent, leading to expanded coil conformations that maximize conformational entropy.
  • In poor solvents:

    • Unfavorable polymer–solvent interactions cause the chain to collapse to minimize the interfacial area with the solvent.
    • This collapse reduces the number of accessible configurations → entropic penalty.
    • Chain becomes confined in a smaller volume, especially under excluded volume constraints.

Entropy Loss Sources:

  • Loss of translational freedom of segments
  • Restricted bond rotations (cis/trans, gauche/trans preferences)
  • Topological entanglement or loop formation in compact states

As solvent quality worsens:

  1. Coil → Globule Transition:

    • Entropy ↓ (chain compaction)
    • Enthalpy ↓ (less polymer–solvent contact)
    • Critical transition when χ exceeds θ-point value in Flory–Huggins theory
  2. Globule → Helix or Aggregated States (in selective systems):

    • Helices form when specific interactions (e.g., hydrogen bonding, stereoelectronic bias) stabilize ordered states
    • Entropy is further reduced, but enthalpic gain (e.g., intrachain H-bonding) can outweigh it
    • Seen in PEG/IBA systems, where helices emerge only under marginal solubility


Flory-type Free Energy (for coil-to-globule transition):

F(R)=(3kBTR^2)/(2Nb^2)+kBT⋅χ⋅(N^2b^6)/R^3

  • First term: entropy of ideal chain (favoring expansion)
  • Second term: enthalpic term from solvent quality (favoring compaction)
  • Minimizing F(R) predicts the transition radius and chain conformation
300

Compare microdiffraction and GIXRD for probing strain localization. When is each technique preferable?


Microdiffraction (μXRD):
– Uses focused X-ray beams (~sub-micron)
– Probes local strain variation within grains or features
– Ideal for mapping strain heterogeneity in complex nanostructures

Grazing-Incidence XRD (GIXRD):
– Uses shallow incidence angles
– Probes surface-near layers (depth-sensitive)
– Best for thin films, surface stress, or layered structures

Use μXRD when spatial resolution is needed (e.g. intra-grain strain).
Use GIXRD when depth profiling or average surface strain is sufficient

300

This technique was used to refine structure factors from SAXS data and confirm unit cell parameters.

Le Bail refinement

300

This combination of mechanical property changes indicated that the 650°C treatment successfully mitigated process-induced anisotropy while enhancing performance.

the simultaneous increase in yield strength and uniform elongation with reduced strength anisotropy

300

These lamellar components persist in solution long after the amorphous phase has degraded, contributing to the long-term environmental presence of nanoplastics.

crystalline lamellae

400

Figure 1: Explain Guinier → Fractal → Porod transitions in the SANS data. What structural features do they indicate about PEG in IBA/H₂O?

PEG in D₂O (open circles):

  • Slope = –1.66 ± 0.13
  • Consistent with a swollen coil conformation
  • Matches the expected exponent for a self-avoiding walk (SAW) in a good solvent → polymer is expanded due to favorable PEG–water interactions

PEG in d-IBA (solid circles):

  • Slope = –0.987 ± 0.007
  • Indicates a rod-like structure → highly anisotropic, stiffened chain
  • Suggests formation of helical or collapsed conformations due to poor solvent conditions and intrachain interactions dominating

PEG in D₂O (good solvent):

  • Strong H-bonding with water
  • Chains adopt entropically favored coil conformations
  • Greater excluded volume → higher radius of gyration
  • Slope near –1.7 aligns with fractal dimension Df≈5/3D_f \approx 5/3Df≈5/3

PEG in d-IBA (poor solvent):

  • Poor solvation causes chain collapse
  • PEG chains adopt rod-like or helical conformations

    • Helices exhibit uniform stiffness and persistence length
    • Slope ~–1 aligns with rigid cylinder (rod) scattering behavior
  • Chain entropy is reduced, but enthalpic stabilization via PEG–PEG contacts (or solvent-mediated ordering) dominates
400

Figure 2 shows the evolution of principal strains and stresses in damascene Cu lines as a function of temperature. Explain the origin of anisotropic strain behavior (ε₁ ≠ ε₂ ≠ ε₃), and describe how thermal expansion mismatch contributes to the observed compressive stress state.


Anisotropic strain (ε₁ ≠ ε₂ ≠ ε₃) arises from confinement of Cu lines in dielectric trenches, leading to directional constraint during thermal expansion.

Thermal expansion mismatch between Cu and surrounding materials (FSG, Si) induces biaxial compressive stress upon heating:
– Cu expands more than FSG/Si
– Lateral expansion (ε₁, ε₂) is constrained
– Vertical direction (ε₃) can partially relax

This mismatch generates triaxial thermoelastic stress, with suppressed plasticity due to stress cancellation in resolved directions.

400

This type of chain-end arrangement leads to six intersecting tubes and is stabilized by weak hydrogen bonding and PEO crystallization

medial packing

400

The increase in strength after 650°C treatment is attributed to this thermally-activated mechanism, which reduces energy barriers for dislocation motion without dissolving cellular walls.

recovery (as opposed to recrystallization)

400

According to this classic statistical model, successive fragmentation of lamellar stacks yields a lateral size distribution consistent with environmental nanoplastic samples.

Kolmogorov log-normal fragmentation model

500

Use partition function arguments to explain helix–coil transitions. What experimental features reflect this?

The helix–coil transition arises from a balance between enthalpic stabilization of helices and entropic favorability of coils in the partition function. Experimentally, it is reflected by a slope change in SANS (–1 to –5/3), p(r) profile shifts, temperature-dependent optical rotation (polarimetry), and a decrease in scattering invariant Q with heating.



500

An X-ray diffraction study shows FWHM variation as a function of ψ angle. Explain how ψ-dependent peak broadening can be used to assess anisotropic strain. What does an increase in FWHM with ψ imply about stress states in thin films?

ψ-dependent peak broadening reflects orientation-specific microstrain. By measuring FWHM at various ψ (tilt) angles, one detects how strain varies with lattice plane orientation.

An increase in FWHM with ψ indicates:
– Greater in-plane strain fluctuations than out-of-plane
– Suggests biaxial stress state, common in thin films
– Points to anisotropic defect distributions or interface-induced strain gradients

This analysis helps reveal strain anisotropy and gradients, even when average strain is undetectable.

500

Despite representing only ~1% of the chain, this modification dramatically alters copolymer phase behavior.

end group functionalization

500

This heat treatment led to a significant increase in yield strength of the AM stainless steel by reducing dislocation density and promoting stress relaxation without full recrystallization.

650°C heat treatment for 1 hour

500

The degradation-induced loss of mechanical integrity in semicrystalline polymers is predicted using this decay model for tie-molecule density as a function of reactive bond fraction nk∼(1−ϕ)ℓk


logarithmic decay model tied to the average number of cleavable bonds per chain type