buckling-induced kirigami - arxiv · buckling-induced kirigami ahmad rafsanjani 1and katia...

11
Buckling-Induced Kirigami Ahmad Rafsanjani 1 and Katia Bertoldi 1,2* 1 John A. Paulson School of Engineering and Applied Sciences, Harvard University, Cambridge, MA 02138, USA and 2 Kavli Institute, Harvard University, Cambridge, MA 02138, USA (Dated: February 22, 2017) We investigate the mechanical response of thin sheets perforated with a square array of mutually orthogonal cuts, which leaves a network of squares connected by small ligaments. Our combined analytical, experimental and numerical results indicate that under uniaxial tension the ligaments buckle out-of-plane, inducing the formation of 3D patterns whose morphology is controlled by the load direction. We also find that by largely stretching the buckled perforated sheets, plastic strains develop in the ligaments. This gives rise to the formation of kirigami sheets comprising periodic distribution of cuts and permanent folds. As such, the proposed buckling-induced pop-up strategy points to a simple route for manufacturing complex morphable structures out of flat perforated sheets. In recent years, origami [19] and kirigami [1027] have become emergent tools to design programmable and reconfigurable mechanical metamaterials. Origami- inspired metamaterials are created by folding thin sheets along predefined creases, whereas kirigami allows the practitioner to exploit cuts in addition to folds to achieve large deformations and create 3D objects from a flat sheet. Therefore, kirigami principles have been exploited to design highly stretchable devices [1824] and morphable structures [2527]. Interestingly, several of these studies also show that pre-creased folds are not necessary to form complex 3D patterns, as mechanical instabilities in flat sheets with an embedded array of cuts can result in out-of-plane deformation [1926]. However, while a wide range of 3D architectures have been realized by triggering buckling under compressive stresses [25, 26], instability-induced kirigami designs subjected to tensile loading are limited to a single incision pattern comprised of parallel cuts in a centered rectangular arrangement [1923]. In this Letter, we investigate the tensile response of elastic sheets of thickness t perforated with a square array of mutually orthogonal cuts. This perforation pattern introduces a network of square domains of edge l separated by hinges of width δ [Fig. 1(a)]. While the planar response of such perforated sheets in the thick limit (i.e. for large values of t/δ) has received significant attention, as it is characterized by effective negative Poisson’s ratio [2836] [Fig. 1(b)], here we add another dimension and study how the behavior of the system evolves when the thickness is progressively decreased (i.e. for decreasing values of t/δ). Our combined analytical, numerical and experimental results indicate that in sufficiently thin sheets mechanical instabilities triggered under uniaxial tension can be exploited to create complex 3D patterns and even to guide the formation of permanent folds. We also find that the morphology of the instability-induced patterns is strongly affected by the loading direction (see Figs. 1(c)- (d) and Movies 1 in Supplemental Material [37]), pointing to an effective strategy to realize functional surfaces characterized by a variety of architectures. We start by experimentally investigating the effect of the sheet thickness t and hinge width δ on the response of the system subjected to uniaxial tension along the square diagonals [i.e. for γ = 45 - Fig. 1(c)]. Specimens are fabricated by laser cutting an array of 3 × 8 mutually perpendicular cuts [see Fig. 2(b)] into plastic sheets (Artus Corporation, NJ) with Young’s modulus E = 4.33 GPa and Poisson’s ratio ν 0.4 (see Supplemental Material: Experiments [37]). In Fig. 2(a), we report the experimental stress-strain responses for 10 samples characterized by different values of normalized thickness t/δ and normalized hinge width δ/l. (a) (b) (c) (d) FIG. 1: (a) Schematic of the system: an elastic sheet of thickness t perforated with a square array of mutually orthogonal cuts. (b) In the thick limit (i.e. for large values of t/δ) the perforated sheet deforms in plane and identically to a network of rotating squares [28]. (c-d) For sufficiently small values of t/δ mechanical instabilities triggered under uniaxial tension result in the formation of complex 3D patterns, which are affected by the loading direction. The 3D patterns obtained for γ = 45 and γ =0 are shown in (c) and (d), respectively. Scale bars: 6 mm. arXiv:1702.06470v1 [cond-mat.soft] 21 Feb 2017

Upload: trannhan

Post on 22-May-2018

215 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: Buckling-Induced Kirigami - arXiv · Buckling-Induced Kirigami Ahmad Rafsanjani 1and Katia Bertoldi;2 1John A. Paulson School of Engineering and Applied Sciences, Harvard University,

Buckling-Induced Kirigami

Ahmad Rafsanjani1 and Katia Bertoldi1,2∗1John A. Paulson School of Engineering and Applied Sciences,

Harvard University, Cambridge, MA 02138, USA and2Kavli Institute, Harvard University, Cambridge, MA 02138, USA

(Dated: February 22, 2017)

We investigate the mechanical response of thin sheets perforated with a square array of mutuallyorthogonal cuts, which leaves a network of squares connected by small ligaments. Our combinedanalytical, experimental and numerical results indicate that under uniaxial tension the ligamentsbuckle out-of-plane, inducing the formation of 3D patterns whose morphology is controlled by theload direction. We also find that by largely stretching the buckled perforated sheets, plastic strainsdevelop in the ligaments. This gives rise to the formation of kirigami sheets comprising periodicdistribution of cuts and permanent folds. As such, the proposed buckling-induced pop-up strategypoints to a simple route for manufacturing complex morphable structures out of flat perforatedsheets.

In recent years, origami [1–9] and kirigami [10–27]have become emergent tools to design programmableand reconfigurable mechanical metamaterials. Origami-inspired metamaterials are created by folding thin sheetsalong predefined creases, whereas kirigami allows thepractitioner to exploit cuts in addition to folds toachieve large deformations and create 3D objects froma flat sheet. Therefore, kirigami principles have beenexploited to design highly stretchable devices [18–24] andmorphable structures [25–27]. Interestingly, several ofthese studies also show that pre-creased folds are notnecessary to form complex 3D patterns, as mechanicalinstabilities in flat sheets with an embedded array ofcuts can result in out-of-plane deformation [19–26].However, while a wide range of 3D architectures havebeen realized by triggering buckling under compressivestresses [25, 26], instability-induced kirigami designssubjected to tensile loading are limited to a singleincision pattern comprised of parallel cuts in a centeredrectangular arrangement [19–23].

In this Letter, we investigate the tensile response ofelastic sheets of thickness t perforated with a squarearray of mutually orthogonal cuts. This perforationpattern introduces a network of square domains of edgel separated by hinges of width δ [Fig. 1(a)]. While theplanar response of such perforated sheets in the thicklimit (i.e. for large values of t/δ) has received significantattention, as it is characterized by effective negativePoisson’s ratio [28–36] [Fig. 1(b)], here we add anotherdimension and study how the behavior of the systemevolves when the thickness is progressively decreased(i.e. for decreasing values of t/δ). Our combinedanalytical, numerical and experimental results indicatethat in sufficiently thin sheets mechanical instabilitiestriggered under uniaxial tension can be exploited tocreate complex 3D patterns and even to guide theformation of permanent folds. We also find thatthe morphology of the instability-induced patterns isstrongly affected by the loading direction (see Figs. 1(c)-

(d) and Movies 1 in Supplemental Material [37]), pointingto an effective strategy to realize functional surfacescharacterized by a variety of architectures.

We start by experimentally investigating the effect ofthe sheet thickness t and hinge width δ on the response ofthe system subjected to uniaxial tension along the squarediagonals [i.e. for γ = 45◦ - Fig. 1(c)]. Specimens arefabricated by laser cutting an array of 3 × 8 mutuallyperpendicular cuts [see Fig. 2(b)] into plastic sheets(Artus Corporation, NJ) with Young’s modulus E =4.33 GPa and Poisson’s ratio ν ' 0.4 (see SupplementalMaterial: Experiments [37]). In Fig. 2(a), we reportthe experimental stress-strain responses for 10 samplescharacterized by different values of normalized thicknesst/δ and normalized hinge width δ/l.

(a)

(b)

(c) (d)

FIG. 1: (a) Schematic of the system: an elastic sheetof thickness t perforated with a square array of mutuallyorthogonal cuts. (b) In the thick limit (i.e. for large values oft/δ) the perforated sheet deforms in plane and identically toa network of rotating squares [28]. (c-d) For sufficiently smallvalues of t/δ mechanical instabilities triggered under uniaxialtension result in the formation of complex 3D patterns, whichare affected by the loading direction. The 3D patternsobtained for γ = 45◦ and γ = 0◦ are shown in (c) and (d),respectively. Scale bars: 6 mm.

arX

iv:1

702.

0647

0v1

[co

nd-m

at.s

oft]

21

Feb

2017

Page 2: Buckling-Induced Kirigami - arXiv · Buckling-Induced Kirigami Ahmad Rafsanjani 1and Katia Bertoldi;2 1John A. Paulson School of Engineering and Applied Sciences, Harvard University,

2

(a)

0 0.05 0.1 0.15 0.2 0.250

0.01

0.02

0.03

0.04

0.05

the

ory

pla

na

r a

uxe

tic (c)(b)

(0.06, 0.085)

(0.06, 0.125)

(0.06, 0.210)

(0.10, 0.051)

(0.10, 0.076)

(0.10, 0.127)

(0.14, 0.036)

(0.14, 0.054)

(0.14, 0.091)

(0.10, 0.508)

0 0.01 0.02 0.03 0.04 0.050

0.005

0.01

0.015

0.02

0.025

theory

FIG. 2: (a) Experimental stress-strain curves for perforated sheets characterized by different normalized hinge width δ/land normalized sheet thickness t/δ for γ = 45◦. Note that the stress is normalized by the effective in-plane Young’s modulusE = 2/3E(δ/l)2. (b) Snapshots of the sample with δ/l = 0.06 and t/δ ' 0.085 at εx = 0, 0.12 and 0.24. (c) Critical strain εcas a function of (t/δ)2 as obtained from experiments (markers) and predicted analytically (dashed line).

First, it is apparent that the initial response for allsamples is linear. At this stage, all hinges bend in-plane,inducing pronounced rotations of the square domains[Fig. 1(b)], which result in large negative values ofthe macroscopic Poisson’s ratio [29, 30]. As such, thestiffness of the perforated sheets, E, is governed by thein-plane flexural deformation of the hinges and it canbe shown that (see Supplemental Material: AnalyticalExploration [37]):

E =σxεx

=2

3E

l

)2

. (1)

Second, for the thin samples (i.e. t/δ � 1), the curvesreported in Fig. 2(a) also show a sudden departure fromlinearity to a plateau stress caused by the out-of-planebuckling of the hinges. Such buckling in turn inducesout-of plane rotations of both the square domains and thecuts, which arrange to form a 3D pattern reminiscent ofa misaligned Miura-ori [38] with an alternation of squaresolid faces (corresponding to the square domains) andrhombic open ones (defined by the cuts) (see Fig. 1(c),Fig. 2(b) at εx = 0.12 and Movie 2 in SupplementalMaterial [37]). To characterize the critical strain, εc, atwhich the instability is triggered, we start by noting thatsince the stress immediately after instability is almostconstant, the contribution of out-of-plane strain energyUo should be linear in εx, (see Supplemental Material:Analytical exploration [37])

Uo(εx) = Eεc(εx − εc). (2)

Moreover, assuming that the square domains remain rigidand that the deformation localizes at the hinges whichcan be modeled as flexural beam segments, Uo can alsobe written as

Uo(εx) = 8× 1

8l2t

∫ δ

0

EIoρ2o

ds =1

3E

(t

l

)2

θ2o, (3)

where Io = δt3/12, ρo = δ/2θo and 2θo is the openingangle of each cut after out-of-plane buckling, which forγ = 450 is approximated by

θ2o ' εx − εc. (4)

Finally, by equating Eqs. (2) and (3) we find that

εc '1

2

(t

δ

)2

, (5)

which despite the simplifications made, compares verywell with our experimental results [Fig. 2(c)] andnumerical simulations [Fig. S6]. Note that a similarexpression for the critical strain has been previouslyobtained for kirigami patterns comprising parallel cutsin a centered rectangular arrangement [23].

(a) (b)

(c) (d)

FIG. 3: The buckling-induced Miura kirigami sheet (a) is flat-foldable, (b) forms a saddle shape with a negative Gaussiancurvature upon non-planar bending, (c) twists under anti-symmetric out-of-plane deformation (d) has much higherbending rigidity than the corresponding flat perforated sheet(inset). Note that the 127 µm thick Miura kirigami sheetshown here supports a 20g weight.

Page 3: Buckling-Induced Kirigami - arXiv · Buckling-Induced Kirigami Ahmad Rafsanjani 1and Katia Bertoldi;2 1John A. Paulson School of Engineering and Applied Sciences, Harvard University,

3

(a) (b)

0 0.05 0.1 0.15 0.2 0.250

0.01

0.02

0.03

0.04

0.05

0 0.05 0.1 0.15 0.2 0.25

-1

-0.5

0

0.5

1

(d)0.8

0.6

0.4

0.2

0

(c)

0 0.05 0.1 0.15 0.20

10

20

30

40

50

0.25

FIG. 4: Effect of loading direction γ on mechanical response of the perforated sheets. Evolution of the (a) normalized stressσx/E, (b) the in-plane macroscopic Poisson’s ratio νyx and (c) the opening angle of cuts 2θo1 and 2θo2 as a function of theapplied strain εx for different values of γ. Note that νyx is negative only for εx < εc (as at this stage the deformation of thestructure is purely planar and identical to that of a network of rotating squares) and that it increases sharply and reachespositive values once the instability is triggered. (d) Numerical snapshots of 3D patterns obtained at εx = 0.125 for differentvalues of γ. The contours shows the normalized out-of-plane displacements.

Third, for large enough values of the applied strainεx, the stress σx rises sharply again. This regimestarts when the square domains align [Fig. 2(b) atεx = 0.24] and the deformation mechanism of the hingesswitches from bending- to stretching-dominated. Atthis stage, localized zone of intense strain (of plasticnature) develop in the hinges and result in the formationof permanent folds. Although we start with a flatelastic sheet with an embedded array of cuts (i.e. aperforated sheet), by largely stretching it we form asystem that comprises a periodic distribution of bothcuts and folds (i.e. a kirigami sheet). In particular, wenote that our kirigami sheets possess several deformationcharacteristics of the Miura-ori [2, 3] and zigzag-basefolded kirigami [12, 13] (see Movie 3 in SupplementalMaterial [37]), as (i) they are flat-foldable [Fig. 3(a)];(ii) they form a saddle shape with a negative Gaussiancurvature upon non-planar bending [Fig. 3(b)]; and(iii) they can be twisted under anti-symmetric out-of-plane deformation, [Fig. 3(c)]. However, in contrastto the Miura-ori, misaligned Miura-ori and zigzag-basefolded kirigami, the macroscopic Poisson’s ratio of ourkirigami sheets is positive (see Movie 4 in SupplementalMaterial [37]). This is the result of the fact that notall the faces are rigid. As such, the applied tensiledeformation not only results in the rotation of the facesabout the connecting ridges, but also in the deformationof those defined by the cuts, allowing lateral contractionof the structure. It is also noteworthy that, differently

from the misaligned Miura-ori that can only be foldedto a plane, the additional degree of freedom provided bythe open cuts allow the Miura kirigami to be laterallyflat-foldable [Movie 4]. Finally, we note that our Miurakirigami structures have higher bending rigidity than thecorresponding flat perforated sheet (see Fig. 3(d) andMovie 3 in Supplemental Material [37]).

Having determined that instabilities in thin sheetswith an embedded array of mutually perpendicularcuts can be harnessed to form complex 3D patterns,we further explore the design space using FiniteElement (FE) analyses (See Supplemental Material: FESimulations [37]). We start by numerically investigatingthe response of finite size samples stretched along thesquare diagonals (i.e. γ = 45◦) and find excellentagreement with the experimental results (Fig. S5 andMovie 2 in Supplemental Material [37]). This validatesthe numerical analyses and indicates that they canbe effectively used to explore the response of thesystem. First, we use the simulations to understandhow plastic deformation evolves. By monitoringthe distribution of the von Mises stress within thesheets, we find that that plastic deformation initiatesat the tip of hinges well after the buckling onset[see Figs. S6 and S9] and then gradually expandto fully cover the hinges when the sample is fullystretched and the deformation mechanism changes frombending-dominated to stretching-dominated. Second,we numerically explore the effect of different loading

Page 4: Buckling-Induced Kirigami - arXiv · Buckling-Induced Kirigami Ahmad Rafsanjani 1and Katia Bertoldi;2 1John A. Paulson School of Engineering and Applied Sciences, Harvard University,

4

conditions and find that uniaxial tension is the idealone to trigger the formation of well-organized out-of-plane patterns in our perforated sheets [see Fig.S7]. Third, we investigate the effect of the loadingdirection by simulating the response of periodic unitcells. In Fig. 4(a) we report the stress-strain responsesobtained numerically for perforated sheets characterizedby t/δ = 0.127 and δ/l = 0.04 loaded uniaxially forγ = 0◦, 15◦, 30◦ and 45◦. Our results indicate thatthe mechanical response of the perforated sheets underuniaxial tension is minimally affected by the loadingdirection. In fact, the evolution of both stress [Fig. 4(a)]and macroscopic in-plane Poisson’s ratio [Fig. 4(b)] aresimilar for different values of γ. By contrast, we findthat the morphology of the 3D patterns induced bythe instability is significantly affected by γ [Fig. 4(c)-(d)]. As the loading directions varies from γ = 45◦ toγ = 0◦, the symmetry in opening angle of the two setsof perpendicular cuts breaks. While for γ = 45◦ all cutsopen equally (i.e. θo1 = θo2), as we reduce γ, one setbecomes wider (i.e. θo1 monotonically increases) andthe other progressively narrower (i.e. θo2 monotonicallydecreases)[Fig. 4(c)]. In the limit case of γ = 0◦ one setof cuts remains almost closed and a 3D cubic patternemerge after buckling [Fig. 1(d), Movie 5]. Furthermore,permanent folds with direction controlled by γ can beintroduced by largely stretching the perforated sheets.As such, by controlling the loading direction a varietyof kirigami sheets can be formed [Movie 6]. While all ofthem are laterally flat-foldable, we find that by increasingγ from 0◦ to 45◦ the resulting kirigami sheets have higherbending rigidity and their Gaussian curvature varies fromzero (for γ = 0◦) to large negative values (for γ = 45◦).Furthermore, by increasing γ, the resulting kirigamisheets become more compliant under torsion (Movie 6in Supplemental Material [37]).

In summary, our combined experimental, analyticaland numerical study indicates that buckling in thinsheets perforated with a square array of cuts andsubjected to uniaxial tension can be exploited to form3D patterns and even create periodic arrangements ofpermanent folds. While buckling phenomena in crackedthin plates subjected to tension have traditionally beenregarded as a route toward failure [39], we show that theycan also be exploited to transform flat perforated sheetsto kirigami surfaces. Our buckling-induced strategy notonly provides a simple route for manufacturing kirigamisheets, but can also be combined with optimizationtechniques to design perforated patterns capable ofgenerating desired complex 3D surfaces under externalloading [9, 11, 41]. Finally, since the response of ourperforated sheets is essentially scale-free, the proposedpop-up strategy can be used to fabricate kirigami sheetsover a wide range of scales, from transformable meter-scale architectures to tunable nano-scale surfaces [24, 40].

Acknowledgements

K.B. acknowledges support from the National ScienceFoundation under grant number DMR-1420570 andCMMI-1149456. A.R. also acknowledges the financialsupport provided by Swiss National Science Foundation(SNSF) under grant number 164648. The authors thankBolei Deng for fruitful discussions, Yuerou Zhang forassistance in laser cutting and Matheus Fernandes forproofreading the manuscript.

∗ Correspondence to [email protected][1] L. Mahadevan and S. Rica, Self-organized origami,

Science 307, 1740 (2005).[2] M. Schenk, and S. D. Guest, Geometry of Miura-folded

metamaterials. Proc. Natl. Acad. Sci. USA 110, 3276(2013).

[3] Z. Y. Wei, Z. V. Guo, L. Dudte, H. Y. Liang, and L.Mahadevan, Geometric Mechanics of Periodic PleatedOrigami, Phys. Rev. Lett. 110, 215501 (2013).

[4] T. Tachi, Designing freeform origami tessellations bygeneralizing Resch’s patterns, ASME J. Mech. Design135, 111006 (2013).

[5] J. L. Silverberg, A. A. Evans, L. McLeod, R. C. Hayward,T. Hull, C. D. Santangelo, and I. Cohen, Using origamidesign principles to fold reprogrammable mechanicalmetamaterials, Science 345, 647 (2014).

[6] H. Yasuda, and J. Yang, Reentrant origami-based meta-materials with negative Poisson’s ratio and bistability.Phys. Rev. Lett. 114, 185502 (2015).

[7] J. L. Silverberg, J-H. Na, A. A. Evans, B. Liu, T. C.Hull, C. D. Santangelo, R. J. Lang, R. C. Hayward, andI. Cohen, Origami structures with a critical transition tobistability arising from hidden degrees of freedom, Nat.Mat. 14, 389 (2015).

[8] J. T. B. Overvelde, T. A. de Jong, Y. Shevchenko,S. A. Becerra, G. M. Whitesides, J. C. Weaver, C.Hoberman, K. Bertoldi, A three-dimensional actuatedorigami-inspired transformable metamaterial with mul-tiple degrees of freedom, Nat. Comm. 7, 10929 (2016).

[9] L. H. Dudte, E. Vouga, T. Tachi, and L. Mahadevan,Programming curvature using origami tessellations, Nat.Mat. 15, 583 (2016).

[10] T. Castle, Y. Cho, X. Gong, E. Jung, D. M. Sussman,S. Yang, and R. D. Kamien, Making the Cut: LatticeKirigami Rules. Phys. Rev. Lett. 113, 245502 (2014).

[11] D. M. Sussman, Y. Cho, T. Castle, X. Gong, E. Jung, S.Yang, and R. D. Kamien, Algorithmic lattice kirigami:A route to pluripotent materials, Proc. Natl. Acad. Sci.USA 112, 7449 (2015).

[12] M. Eidini and G. H. Paulino, Unraveling metamaterialproperties in zigzag-base folded sheets, Sci. Adv. 1,e1500224 (2015).

[13] M. Eidini, Zigzag-base folded sheet cellular mechanicalmetamaterials, Ext. Mech. Lett. 6 96 (2016).

[14] T. Castle, D. M. Sussman, M. Tanis and R. D. Kamien,Additive lattice kirigami Sci. Adv. 2, e1601258 (2016).

[15] B. G. Chen, B. Liu, A. A. Evans, J. Paulose, I. Cohen,V. Vitelli, and C. D. Santangelo. Topological Mechanics

Page 5: Buckling-Induced Kirigami - arXiv · Buckling-Induced Kirigami Ahmad Rafsanjani 1and Katia Bertoldi;2 1John A. Paulson School of Engineering and Applied Sciences, Harvard University,

5

of Origami and Kirigami. Phys. Rev. Lett. 116, 135501(2016).

[16] Y. Tang and J. Yin, Design of cut unit geometry inhierarchical kirigami-based auxetic metamaterials forhigh stretchability and compressibility, Ext. Mech. Lett.in press (2016).

[17] K. A. Seffen, k-cones and kirigami metamaterials Phys.Rev. E 94, 033003 (2016).

[18] Z. Song, X. Wang, C. Lv, Y. An, M. Liang, T. Ma, D. He,Y-J. Zheng, S-Q. Huang, H. Yu, and H. Jiang, Kirigami-based stretchable lithium-ion batteries Sci. Rep. 5, 10988(2015).

[19] T. C. Shyu, P. F. Damasceno, P. M. Dodd, A.Lamoureux, L. Xu, M. Shlian, M. Shtein, S. C. Glotzer,and N. A. Kotov, A kirigami approach to engineeringelasticity in nanocomposites through patterned defects,Nat. Mat. 14, 785 (2015).

[20] M. K. Blees, A. W. Barnard, P. A. Rose, S. P. Roberts,K. L. McGill, P. Y. Huang, A. R. Ruyack, J. W. Kevek,B. Kobrin, D. A. Muller, and P. L. McEuen, Graphenekirigami, Nature 524 204 (2015).

[21] A. Lamoureux, K. Lee, M. Shlian, S. R. Forrest, and M.Shtein, Dynamic kirigami structures for integrated solartracking, Nat. Comm. 6, 8092 (2015).

[22] D. Norman, The mounting of single leaf parchment andvellum objects for display and storage, V&A Conserv. J.,9, 10 (1999).

[23] M. Isobe and K. Okumura, Initial rigid response andsoftening transition of highly stretchable kirigami sheetmaterials, Sci. Rep. 6, 24758 (2016).

[24] C. Wu, X. Wang, L. Lin, H. Guo,and Z. L. Wang, Paper-based triboelectric nanogenerators made of stretchableinterlocking kirigami patterns ACS Nano 10, 4652(2016).

[25] Y. Zhang, Z. Yan, K. Nan, D. Xiao, Y. Liu, H. Luan,H. Fu, X. Wang, Q. Yang, J. Wang, W. Ren, H. Si,F. Liu, L. Yang, H. Li, J. Wang, X. Guo, H. Luo, L.Wang, Y. Huang, J. A. Rogers, A mechanically drivenform of Kirigami as a route to 3D mesostructures inmicro/nanomembranes, Proc. Natl. Acad. Sci. USA 112,11757 (2015).

[26] Z. Yan, F. Zhang, J. Wang, F. Liu, X. Guo, K. Nan, Q.Lin, M. Gao, D. Xiao, Y. Shi, Y. Qiu, H. Luan, J. H.Kim, Y. Wang, H. Luo, M. Han, Y. Huang, Y. Zhang, J.A. Rogers, Controlled Mechanical Buckling for Origami-Inspired Construction of 3D Microstructures in AdvancedMaterials Adv. Func. Mater. 26, 2629 (2016).

[27] R. M. Neville, F. Scarpa, and Alberto Pirrera, Shapemorphing Kirigami mechanical metamaterials Sci. Rep.6, 31067 (2016).

[28] J. N. Grima and K. E. Evans, Auxetic behavior fromrotating squares, J. Mat. Sci. Lett. 19, 1563 (2000).

[29] A. A. Vasiliev, S. V. Dmitriev, Y. Ishibashi and T.Shigenari Elastic properties of a two-dimensional modelof crystals containing particles with rotational degrees offreedom. Phys. Rev. B 65, 094101 (2002).

[30] J. N. Grima, A. Alderson and K. E. Evans, Auxeticbehaviour from rotating rigid units phys. stat. sol. (b)242, 561 (2005).

[31] S. Shan S., S. H. Kang, Z. Zhao, L. Fang, and K.Bertoldi, Design of planar isotropic negative Poisson’sratio structures, Ext. Mech. Lett. 4, 96 (2015).

[32] Y. Cho, J-H Shin, A. Costa, T. A. Kim, V. Kunin, J.Li, S. Yeon Lee, S. Yang, H. N. Han, I-S Choi, and

D. J. Srolovitz, Engineering the shape and structure ofmaterials by fractal cut, Proc. Natl. Acad. Sci. USA 111,17390 (2014).

[33] R. Gatt, L. Mizzi, J. I. Azzopardi, K. M. Azzopardi, D.Attard, A. Casha, J. Briffa, J. N. Grima, Hierarchicalauxetic mechanical metamaterials, Sci. Rep. 5, 8395(2015).

[34] Y. Suzuki, G. Cardone, D. Restrepo, P. D. Zavattieri,T. S. Baker, A. F. Tezcan, Self-assembly of coherentlydynamic, auxetic, two-dimensional protein crystals,Nature 533, 369 (2016).

[35] J. N. Grima, E. Manicaro, and D. Attard, Auxeticbehaviour from connected different-sized squares andrectangles, Proc. R. Soc. A 467, 439 (2011)

[36] A. Rafsanjani and D. Pasini, Bistable Auxetic Mechani-cal Metamaterials Inspired by Ancient Geometric Motifs,Ext. Mech. Lett. 9, 291 (2016).

[37] See Supplemental Material at URL which includesRefs. [28–30], supporting movies, an analytical explo-ration, details of the FE simulations and a descriptionof the experiments.

[38] K. Saito, A. Tsukahara, and Y. Okabe, Designingof self-deploying origami structures using geometricallymisaligned crease patterns, Proc. R. Soc. A 472,20150235 (2016)

[39] G.F. Zielsdorff and R. L. Carlson, On the buckling of thintensioned sheets with cracks and slots, Eng. Frac. Mech.4, 939 (1972).

[40] F. Cavallo, Y. Huang, E. W. Dent, J. C. Williams, andM. G. Lagally, Neurite Guidance and Three-DimensionalConfinement via Compliant Semiconductor Scaffolds,ACS Nano 8, 12219 (2014).

[41] M. Konakovic, K. Crane, B. Deng, S. Bouaziz, D.Piker, and M. Pauly, Beyond developable: computationaldesign and fabrication with auxetic materials ACM.Trans. Graph. 35(4) no. 89 (2016)

Page 6: Buckling-Induced Kirigami - arXiv · Buckling-Induced Kirigami Ahmad Rafsanjani 1and Katia Bertoldi;2 1John A. Paulson School of Engineering and Applied Sciences, Harvard University,

6

SUPPLEMENTAL MATERIALS

Analytical Exploration

To get a deeper understanding of the mechanical re-sponse of the considered patterned sheets, we analyticallyinvestigate their behavior. We first study the initialin-plane linear elastic response of the system and thencharacterize the onset of instability resulting in theformation of 3D patterns. In all our calculations weassume that all deformation is localized at the hingesand that the square domains are rigid.

Initial in-plane linear elastic response. Thestress-strain curves shown in Fig. 2a of the main textshow that the response of all samples is initially linear.Here, we derive an analytical relation for the effectiveYoung’s modulus of the perforated sheets, E, in termsof the geometrical parameters l and δ, and the Young’smodulus of the sheet E.

We focus on a unit cell comprising four square domains[Fig. S1(a)] and deform it uniaxially along one set ofcuts (i.e. γ = 0◦) by applying a macroscopic stressσx = f/(2lt), where f is the force applied to thehinges on the vertical boundaries and 2lt denotes itscross sectional area [Fig. S1(b)]. It is important to notethat using standard axis transformation techniques it hasbeen shown that the planar response of such perforatedsheets is not affected by the loading direction γ [29, 30].Therefore, although here for the sake of simplicity weconsider γ = 0◦, we expect E to be identical for anyloading direction (i.e. for any value of γ).

The applied uniaxial stress generates identical bendingmoments M at all hinges, which in turn induce therotation of all square domains by an angle θi and theopening of the all cuts by an angle 2θi [Fig. S1(b)].Focusing on a single square domain [Fig. S1(c)], it is easyto see that:

M =1

4× fl

2=

1

4σxl

2t. (S1)

Moreover, since M must be balanced by the couple

(a) (b) (c)

FIG. S1: (a) Schematic of the system in the undeformedconfiguration: an elastic sheet of thickness t perforated witha square array of mutually orthogonal cuts. (b) Schematicof the system in the planarly deformed configuration (c) Freebody diagram of a square domain.

induced by the internal loads, we have:

M =EIiρi

, (S2)

where Ii = δ3t/12 is the second moment of area of eachhinge about the z-axis and ρi denotes the curvature ofeach bent hinge. Assuming that the length of the bentregion of the hinge is approximately equal to the hingewidth δ, we obtain:

ρi 'δ

2θi(S3)

Substitution of Eqs. (S1) and (S3) into Eq. (S2) yields:

σx =2

3E

l

)2

θi (S4)

Moreover, since the strain in the loading direction isgiven by:

εx = cos θi + sin θi − 1, (S5)

in the small deformation regime (i.e. θi → 0) we have:

εx ' θi, (S6)

so that

σx =2

3E

l

)2

εx (S7)

It follows that the effective Young’s modulus ofperforated sheet, E, is:

E =σxεx

=2

3E

l

)2

. (S8)

Finally, we note that considering each hinge as a beamof thickness δ and width t, its strain energy density underin-plane deformation can be calculated as:

Uhinge =1

2V

∫ δ

0

EIiρ2i

ds =1

16Eε2x (S9)

where V = 4l2t is the volume of the unit cell. The in-plane strain energy density of a unit cell comprising eighthinges is then given by:

Ui = 8× Uhinge =1

2Eε2x (S10)

Transition from in-plane to out-of-plane re-sponse. The stress-strain curves reported in Fig. 1a ofthe main text also show that the thin samples (i.e. t/δ �1) are characterized by a sudden departure from linearityto a plateau stress. This sudden departure from linearityis the result of out-of-plane buckling of the ligamentsand occurs when the out-of-plane deformation of the

Page 7: Buckling-Induced Kirigami - arXiv · Buckling-Induced Kirigami Ahmad Rafsanjani 1and Katia Bertoldi;2 1John A. Paulson School of Engineering and Applied Sciences, Harvard University,

7

alignment of hinges

along the loading direction

in-p

lane d

eform

a�on

out-of-plane deforma�on

in-plane deforma�on

out-of-plane deforma�on

(b)

(c)

(d)

(a)

(e)A

B

FIG. S2: (a) Schematic of the stress-strain relationship for aperforated sheet that deform in-plane (red line) and out-of-plane (red line). (b) Schematic of the evolution of the strainenergy versu the applied strain for a perforated sheet thatdeform in-plane (red line) and out-of-plane (red line). (c)Schematic of a unit cell deforming in-plane. (d) Schematicof a unit cell deforming out-of-plane. (e) Paper model of theunit cell deformed out-of-plane. The model clearly shows thatthe hinges are aligned along the loading direction.

hinges becomes energetically less costly than their in-plane deformation. While the strain energy density of aperforated sheet that has deformed in-plane is given byEq. (S10), the strain energy density of a perforated sheetthat has deformed out-of plane can be expressed as:

Ubuckled(εx) = Ui(εc) + Uo(εx) (S11)

where εc is the critical strain at which buckling occursand Uo is energy contribution due to the out-of-planedeformation.

Since σx = dUbuckled/dεx and the stress immediatelyafter instability is almost constant [Fig. S2(b)], it followsthat Uo(εx) is approximately linear in εx [Fig. S2(a)] andcan be identified as the tangent of Ui(εx) at εc:

Uo(εx) = Eεc(εx − εc) (S12)

Moreover, as all hinges bend out-of-plane, Uo(εx) canalso be expressed as:

Uo(εx) = 4× 1

2V

∫ δ

0

EIoρ2o1

ds+ 4× 1

2V

∫ δ

0

EIoρ2o2

ds =1

6E

(t

l

)2

(θ2o1 + θ2o2) (S13)

where Io = δt3/12 is the second moment of area withrespect to x-axis (y-axis) for hinges along y-axis (x-axis), ρoj = δ/(2θoj ) is the out-of-plane curvature ofthe deformed hinges and 2θoj is the opening angle of anhinge after out-of-plane buckling. Note that, in general,after buckling the opening angles of the hinges within theperforate sheet can take two values, 2θo1 and 2θo2 (i.e.not all hinges open equally after buckling).

While Eq. (S13) is valid for any loading direction,for γ = 45o (i.e. loading along the square diagonals)θo1 = θo2 = θo, so that Eq. (S13) simplifies to:

Uo(εx) = 8× 1

2V

∫ δ

0

EIoρ2o

ds =1

3E

(t

l

)2

θ2o. (S14)

The critical strain εc can then be determined by equationEqs. (S12) and (S14) after having expressed θo as afunction of εx. To this end, we first note that for γ = 45o

[Fig. S2(d)]

cos 2θo = cosα sin 2β (S15)

where α ∈ [0, π/2] determines the orientation of thesquare domains with respect to xy-plane and

β = arccos1 + εx√

2. (S16)

is the angle between the edge of the square domain placedon the xy-plane and the x-axis.

Moreover, since the experiments indicate that forγ = 45◦ the hinges are approximately aligned along theloading direction [see paper illustration in Fig. S2(e)], thedistance between points A and B shown in Fig. S2(d)remains constant (i.e. |AB| =

√5l) and

cosα = tanβ. (S17)

Page 8: Buckling-Induced Kirigami - arXiv · Buckling-Induced Kirigami Ahmad Rafsanjani 1and Katia Bertoldi;2 1John A. Paulson School of Engineering and Applied Sciences, Harvard University,

8

While Eq. (S17) is exact well into the postbucklingregime, it does not correctly capture the response of thesystem at the onset of instability, as it predicts α 6= 0for εc (i.e. it predicts that the square are already rotatedout of the xy-plane when the instability is triggered). Tocorrect for this, we modify Eq. (S17) as

cosα ' tan(β + θc), for εx ≥ εc. (S18)

where θc = π/4−βc denotes the opening angle associatedto εc. Note that, according to Eq. (S18), α = 0 at εcSubstitution of Eqs. (S16) and (S18) into Eq. (S15)yields:

θo(εx) =1

2arccos

[sin

(2 arccos

1 + εx√2

)tan

4+ arccos

1 + εx√2− arccos

1 + εc√2

)](S19)

Although Eq. (S19) provides a highly non-linear relationbetween θo and εx, close to the instability point θ2o(εx)can be approximated as

θ2o(εx) ' εx − εc (S20)

and can be then inserted into Eq. (S11) to express Uoin terms of εx. Finally, the critical strain εc can bedetermined by substituting Eq. (S20) into Eq. (S14) andthen equating it to Eq. (S12):

εc '1

2

(t

δ

)2

(S21)

This relation shows that the critical strain scalesquadratically with t/δ and despite several simplificationsmade, it compares very well with both experimental[Fig. 2(c)] and numerical [Fig. S8] results.

Experiments

Fabrication. Specimens are fabricated by lasercutting an array of 3×8 mutually perpendicular cuts intoplastic sheets (Artus Corporation, NJ). Total number of10 samples are fabricated with a combination of differentsheet thickness (t = 50.8µm, 76.2µm, 127µm and 508µm)and three normalized hinge widths (δ/l = 0.06, 0.1, 0.14).Note that in all our samples l = 10mm. The materialproperties of the plastic sheets used in this study arecharacterized by performing uniaxial tensile tests (ASTMD882) with a uniaxial testing machine (Instron 5566)equipped with a 100N load cell. Strips with a width of10 mm and gauge length of 120 mm are fully clamped atboth ends using a pneumatic gripper and stretched witha displacement rate of 0.1 mm/s up to ε = 0.8 [Fig. S3].The stress-strain curves reported in Fig. S3 indicate thatthe polyester sheets with thickness t = 50.8µm (red line),76.2µm (green line), 127µm (blue line) are characterizedby a Young’s modulus E = 4.33 GPa. Moreover, their

0.2% offset yield strength is measured as σy = 66.4 MPaand their plastic strain εp versus σ is reported in theTable S1 up to fully plastic region. Differently, for thethick PETG sheet (t = 508µm, yellow line) we measurea Young’s modulus E = 1.75 GPa. Finally, we note thatfor all the plastic sheets a typical Poisson’s ratio ν = 0.4is assumed.

εp 0 0.002 0.004 0.006 0.008 0.018 0.028

σ (MPa) 66.4 75.78 80.52 82.75 84.45 88.09 89.06

TABLE S1: Plastic strain (εp) versus stress (σ) for the threepolyester sheets.

Testing. The quasi-static uniaxial tensile response ofthe perforated sheets comprised of 3 × 8 unit cells wasprobed by using a uniaxial testing machine (Instron 5566)equipped with a 10N load cell. All test were conductedunder displacement rate of ux = 0.1 mm/s [Fig. S4].

0 0.02 0.04 0.06 0.080

25

50

75

100

FIG. S3: Stress-strain curves of the plastic sheets obtainedby stretching them uniaxially. The sheets with t = 50.8,76.2 and 127µm are made of polyester, while the one witht = 508µm is made of PETG.

Page 9: Buckling-Induced Kirigami - arXiv · Buckling-Induced Kirigami Ahmad Rafsanjani 1and Katia Bertoldi;2 1John A. Paulson School of Engineering and Applied Sciences, Harvard University,

9

kirigami

sample

pneuma�c

gripper

load cell

pneuma�c

gripper

(a) (b)

FIG. S4: (a) Experiment setup and (b) closeup view of theperforated sheet.

Finite Element simulations

In this Section, we provide details about the FiniteElement (FE) simulations conducted for this study usingthe commercial package Abaqus\Standard 6.12 (DassaultSystemes). In all simulations, the models are discretizedwith 3D shell elements (S4R) and the cuts in flat sheetsare modeled as seam cracks with duplicate overlappingnodes along the cuts.

Finite size simulations. To validate the FEsimulations, we first performed finite size simulations onperforated sheets similar to those used in experimentscomprising an array of 3 × 8 cells. The lower boundaryof the sample is fixed and a vertical displacementis applied to the upper boundary while the lateralboundaries are assumed to be traction free. Thematerial behavior of the plastic sheet is captured usingan elasto-plastic model (material models *ELASTICand *PLASTIC in Abaqus) with the experimentallycharacterized properties (see Fig. S3 and Table S1).The response of the sheets is then simulated conductingdynamic implicit simulations (*DYNAMIC module inAbaqus). To facilitate convergence, we introduce someartificial, numerical damping (by setting the parametersα = −0.41421, β = 0.5 and γ = 0.91421 in the Hilber-Hughes-Taylor time integration algorithm). Moreover,quasi-static conditions are ensured by monitoring thekinetic energy and finally, to trigger the instability animperfection is introduced by applying two opposingsmall bias forces normal to the sheet plane at two endsof each cut during the initial phase of each simulation.

First, we numerically investigate the response offinite size samples stretched along the square diagonals(i.e. γ = 45◦) and find excellent agreement with theexperimental results (Fig. S5 and Movie 2). Thisvalidates the numerical analyses and indicates that theycan be effectively used to explore the response of thesystem. Moreover, the simulations provide additionalinsights, since they allow us to easily monitor the

0 5 10 15 20 25 300

2

4

6

8

10

12

0 5 10 15 20 25 300

2

4

6

8

10

12

Exp FEM Exp FEM

0 5 10 15 20 25 300

1

2

3

4

5

6FEM Exp.

50.8

76.2

127

(d)

plateau regime hardening regime

(a) (b)

(c)

FIG. S5: Comparison between experimental and numericalforce displacement curves for (a) δ/l = 0.06, (b) δ/l = 0.1and (c) δ/l = 0.14 and different thickness t where l = 10mm.(d) Snapshots of experimental and numerical results at εx =0.125.

stress distribution within the sheets and, therefore, tounderstand how plastic deformation evolves. In Fig. S6we show a close-up view of the distribution of von Misesstress, σv, at the hinges located in the middle of a finitesize sample characterized by t/δ = 0.06 (l = 10mmand t = 50.8µm) for γ = 45◦. Since for the materialconsidered in this study plastic deformation developswhen σv = σy = 66.4 MPa (see Fig. S3 and Table S1),the snapshots indicate that yielding at the hinges initiateat εx ' 0.032. Note that, although this is a very smallvalue of strain, it is well beyond the onset of buckling(for this sample εc ' 0.0036). We then find that theplastic zone at the hinges gradually increase with theapplied strain and fully cover them when the sample is

66.4

49.8

33.2

16.6

0

FIG. S6: Von Mises stress distribution σv at the hingeslocated in the middle of a finite size sample characterizedby t/δ = 0.06 (l = 10mm and t = 50.8µm) for γ = 45◦.Snapshots at different strain levels are shown. Note thevon Mises stress reaches the yield strength of the material(σy = 66.4 MPa) when εx = 0.032. Gray contours showthe regions where σv is larger than the yield strength σy (i.e.plastic zones).

Page 10: Buckling-Induced Kirigami - arXiv · Buckling-Induced Kirigami Ahmad Rafsanjani 1and Katia Bertoldi;2 1John A. Paulson School of Engineering and Applied Sciences, Harvard University,

10

(a) (b) (c)

FIG. S7: Numerical snapshots of perforated sheets underbiaxial loading: (a) equibiaxial tension at εx = εy = 0.1,(b) equibiaxial compression at εx = εy = −0.1 and (c) biaxialtension/compression at εx = 0.1 and εy = −0.1.

fully stretched and the deformation mechanism of thehinges changes from bending-dominated to stretching-dominated.

Second, we use FE to explore the effect of differentloading conditions. More specifically, while in the maintext we focus exclusively on perforated sheets subjectedto uniaxial tension, here we investigate the response ofa perforated sheet characterized by t/δ = 0.127 andδ/l = 0.04 and under biaxial deformation applied atγ = 45◦. We perform simulations on a finite sizesample comprising 3 × 3 unit cells and consider threeload cases: (i) equibiaxial tension (i.e. εx = εy > 0),(ii) equibiaxial compression (i.e. εx = εy < 0) and (iii)biaxial tension/compression (i.e. εx = −εy > 0). Forall cases appropriate displacements in the x-y plane areapplied to all nodes on the edges of the models, whileconstraining their displacements in z-direction (note thatall rotations are left unset).

Our simulations indicate that under equibiaxialtension the structure remains roughly flat and no out-of-plane pattern emerges [see Fig. S7(a)]. This is because,as indicated by the Poisson’s ratio results reported inFig. 4(b) of the main text, the formation of the out-of-plane pattern is accompanied by lateral contraction andunder equibiaxial tension such contraction is preventedby the tensile stretch applied in the transverse direction.Moreover, as shown in Fig. S7(b), we find that underequibiaxial compression the periodic pattern of cuts doesnot significantly affect the response of the system. Ourperforated sheet behaves similarly to a continuous thinsheet and buckles out of plane to form a dome-like shape.Finally, for the case of biaxial tension/compression oursimulations show that the response of the perforatedsheets is very similar to that observed under uniaxialtension [see Fig. S7(c)]. This is because, differentlyfrom the case of equibiaxial tension, for this loadingcondition the compressive stretch applied in lateraldirection favors the formation of the out-of-plane pattern.However, it is important to note that in the case ofbiaxial tension/compression the response of the sheets

is highly affected by their size. For sheets with largernumber of unit cell aligned in the direction of theapplied compressive force we find that the sheet bucklesglobally to form a wavy pattern. Therefore, this setof simulations indicate that uniaxial tension is theideal loading condition to trigger the formation of well-organized out-of-plane patterns in our perforated sheets.

Unit cell simulations. To reduce the computationalcosts and make sure the response of the system isnot dominated by boundary effects, we investigate theresponse of infinite perforated sheets under periodicboundary conditions. Since here we are mostly interestedin the response of the perforated sheet immediately afterbuckling (i.e. before the plastic deformation takes place),for this set of simulations we use a linear elastic materialmodel (with E = 4.33 GPa and ν = 0.4). All simulationsconsist of two steps: (i) we first use a linear perturbationanalysis (*BUCKLE module in Abaqus) to identify thecritical buckling mode; (ii) we then introduce a smallimperfection (' 0.001l) in the form of the critical modeinto the mesh to guide the post-buckling analysis. Asfor the finite size simulations, for this step we conductdynamic implicit simulations (*DYNAMIC module inAbaqus) and to facilitate convergence, we introduce someartificial, numerical damping.

In Fig. S8 we compare the analytical expressionfor the critical strain [Eq. (S21)] with the numericalpredictions of 54 unit cell simulations characterized byδ/l ∈ [0.05, 0.1] and t/δ ∈ [0.04, 0.24] and γ = 45◦. Wefind an excellent agreement between numerical (markers)and analytical (dashed line) results.

Finally, in Fig. S9 we show the von Mises stressdistribution in a unit cell characterized by δ/l = 0.04and t/δ = 0.127 for different values of γ at εx = 0.02.As expected, we find that the von Mises stress, σv,is maximum at the hinges. Assuming the sheets aremade of the same material used to fabricate our sample(i.e. with E = 4.33 GPa and ν = 0.4), we also findthat in all unit cells max(σv) ∼ 66 MPa. This is the

0 0.01 0.02 0.03 0.04 0.050

0.005

0.01

0.015

0.02

0.025

FIG. S8: Comparison of the critical strain εc as predicted byFE simulations (markers) and theory [dashed line - Eq. (S21)]for γ = 45◦.

Page 11: Buckling-Induced Kirigami - arXiv · Buckling-Induced Kirigami Ahmad Rafsanjani 1and Katia Bertoldi;2 1John A. Paulson School of Engineering and Applied Sciences, Harvard University,

11

66.4

49.8

33.2

16.6

0

FIG. S9: Numerical snapshots showing the distribution ofthe von Mises stress, σv, in periodic unit cells loaded alongdifferent orientations γ at εx = 0.02. we note that at this levelof applied deformation the von Mises stress at the tips of thecuts reach the yield strength of the sheet material (σy = 66.4MPa).

stress at which we expect plastic deformation to initiatefor the considered material (see Fig. S3 and Table S1).Therefore, we can conclude that the yielding of the hingesbegins approximately at εx ' 0.02. Since, this strain ismore than twice the critical strain (i.e. εc ' 0.008),these simulations confirm that no plastic deformationtakes place before buckling.

Movie captions

Movie 1 Buckling-induced 3D kirigami in thinperforated sheets.

In sufficiently thin sheets perforated with a squarearray of mutually orthogonal cuts mechanical instabilitiesare triggered under uniaxial tension and can be exploitedto create complex 3D patterns. If the sheet is loadedalong the square diagonals (i.e. γ = 45◦), a 3D patternreminiscent of a misaligned Miura-ori emerges, whileloading along one set of the cuts (i.e. γ = 0◦) results ina 3D cubic pattern.

Movie 2 Uniaxial loading: Experiment vs FEsimulation.

The deformation of perforated sheets subjected touniaxial tensile loading along the square diagonals

(γ = 45◦) can be accurately captured by FE simulationsin Abaqus.

Movie 3 Buckling-induced Miura kirigami .

Although we start with a flat elastic sheet withan embedded array of cuts, by largely stretching itwe end up with a system that comprises a periodicarrangement of both cuts and folds. As a result,after being fully stretched our sheets possess severaldeformation characteristics of the Miura-ori, includingflat-foldability, negative Gaussian curvature undernon-planar bending and twisting under anti-symmetricout-of-plane deformation. Miura kirigami also exhibitsan enhanced bending rigidity compared to a flatperforated sheet.

Movie 4 In-plane Poisson’s ratio of buckling-inducedkirigami.

In contrast to Miura-ori and misaligned Miura-ori,the in-plane Poissons ratio of our perforated sheets afterbuckling is positive.

Movie 5 Buckling-induced cubic kirigami.

The kirigami structure obtained by fully stretching theperforated sheet along one set of the cuts (i.e. γ = 0◦) is(i) flat foldable, (ii) exhibits a zero Gaussian curvatureunder non-planar bending and (iii) has relatively highertorsional rigidity compared to that of the structureobtained by loading the sheet along the diagonals of thesquares (i.e. γ = 45◦).

Movie 6 Comparing the behavior of kirigami sheets.

Permanent folds with direction controlled by γ can beintroduced by largely stretching the perforated sheets.As such, by controlling the loading direction a varietyof kirigami sheets can be formed. While all of themare laterally flat-foldable, we find that by increasing γfrom 0◦ to 45◦ the resulting kirigami sheets have higherbending rigidity and their Gaussian curvature varies fromzero (for γ = 0◦) to large negative values (for γ = 45◦).Furthermore, by increasing γ, the resulting kirigamisheets become more compliant under torsion.