12
Zhang et al., Sci. Adv. 2021; 7 : eabf1966 24 February 2021 SCIENCE ADVANCES | RESEARCH ARTICLE 1 of 11 MATERIALS SCIENCE Hierarchical mechanical metamaterials built with scalable tristable elements for ternary logic operation and amplitude modulation Hang Zhang 1,2 , Jun Wu 1,2 , Daining Fang 3 *, Yihui Zhang 1,2 * Multistable mechanical metamaterials are artificial materials whose microarchitectures offer more than two dif- ferent stable configurations. Existing multistable mechanical metamaterials mainly rely on origami/kirigami- inspired designs, snap-through instability, and microstructured soft mechanisms, with mostly bistable fundamental unit cells. Scalable, tristable structural elements that can be built up to form mechanical metamaterials with an extremely large number of programmable stable configurations remains illusive. Here, we harness the elastic tensile/ compressive asymmetry of kirigami microstructures to design a class of scalable X-shaped tristable structures. Using these structure as building block elements, hierarchical mechanical metamaterials with one-dimensional (1D) cylindrical geometries, 2D square lattices, and 3D cubic/octahedral lattices are designed and demonstrated, with capabilities of torsional multistability or independent controlled multidirectional multistability. The number of stable states increases exponentially with the cell number of mechanical metamaterials. The versatile multistability and structural diversity allow demonstrative applications in mechanical ternary logic operators and amplitude modulators with unusual functionalities. INTRODUCTION Mechanical metamaterials (116) represent a type of artificial mate- rials usually consisting of periodic microstructures whose architec- tures are carefully designed to offer mechanical properties that surpass those of conventional materials. These architected metama- terials mainly leverage the spatial motions, extreme deformations, multiple equilibrium states, and shape morphing of microstructures to obtain exotic properties and/or functionalities, such as zero/negative values of Poisson’s ratios (1720), thermal expansion coefficients (2124) and swelling ratios (2526), reprogrammable stiffness and/ or dissipation (2733), controlled acoustic wave propagation (3435), and tailorable multistability (93637). The latter property (i.e., tailorable multistability) is of rapidly increasing interests because of promising potentials for applications in information processing (3841), recyclable energy absorption (4243), and soft robotic sys- tems (4447). These materials, sometimes termed as “multistable mechanical metamaterials,” are designed to offer more than two stable states that can be switched reversibly among each other. This requires an elaborate manipulation of the energy landscape, and several strategies have been reported, including those that rely on origami/kirigami-inspired designs (4854), snap-through instability (425559), microstructured soft mechanisms (364360), and geo- metrical frustration (6163). For example, the diversity and high foldability of prismatic geometries have been leveraged in the con- text of origami techniques to develop a class of three-dimensional (3D) multistable metamaterials with periodic arrangements of rigid plates and elastic hinges (495253). The instability-based strategies mainly exploited elastic beams capable of snapping between two different stable configurations to create bistable building block ele- ments that can be further extended to form multistable metamate- rials (40425564). Despite these important progresses, it remains challenging to design hierarchical metamaterials with thousands of stable states and precisely tailored steady-state properties. In partic- ular, ample opportunities exist in the development of scalable, tristable structural elements that can be built up to form multistable mechanical metamaterials with an extremely high number (e.g., >10 4 ) of programmable stable configurations. This paper introduces a class of X-shaped kirigami microstruc- tures as tristable building block elements, which can be extended, following a bottom-up scheme, to achieve hierarchical mechanical metamaterials with an exponentially increased number of stable states. Here, the tristability arises mainly from the elastic tensile/ compressive asymmetry of kirigami microstructures, representing a distinct mechanism from those exploited in previous designs of multistable mechanical metamaterials (36374042435561). Multimaterial 3D printing technologies enable the fabrication and experimental validation of the programmable multistability in me- chanical metamaterials with 1D cylindrical geometries, 2D square lattices, and 3D cubic/octahedral lattices. The number of stable states in the developed 2D multistable mechanical metamaterials with (M × N) unit cells follows an exponential law of (3 M + N ) due to the independently controlled multistability along diverse directions. Note that although a few tristable structures have been reported (6567), they are not scalable and cannot serve as building block elements of multistable mechanical metamaterials. Quantitative studies of the underlying mechanics establish the relationship between the geom- etries of X-shaped kirigami microstructures and the resulting energy landscape. The unique tristable building block structure allows the design and demonstration of fundamental mechanical ternary logic gates, as well as complex logic operators, which are unachievable previously. Compared with binary logic gates, much more informa- tion can be processed through ternary/multivalued logic gates, which can effectively reduce the design complexity and speed up the serial 1 AML, Department of Engineering Mechanics, Tsinghua University, Beijing 100084, P.R. China. 2 Center for Flexible Electronics Technology, Tsinghua University, Beijing 100084, P.R. China. 3 Institute of Advanced Structure Technology, Beijing Key Labo- ratory of Lightweight Multi-Functional Composite Materials and Structures, Beijing Institute of Technology, Beijing 100081, P.R. China. *Corresponding author. Email: [email protected] (D.F.); yihuizhang@tsinghua. edu.cn (Y.Z.) Copyright © 2021 The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original U.S. Government Works. Distributed under a Creative Commons Attribution NonCommercial License 4.0 (CC BY-NC). on August 11, 2021 http://advances.sciencemag.org/ Downloaded from

MATERIALS SCIENCE Copyright © 2021 Hierarchical mechanical … · Zhang et al., Sci. dv. 2021 7 : eabf1966 24 February 2021 SCIENCE ADANCES | RESEARCH ARTICLE 1 of 11 MATERIALS SCIENCE

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: MATERIALS SCIENCE Copyright © 2021 Hierarchical mechanical … · Zhang et al., Sci. dv. 2021 7 : eabf1966 24 February 2021 SCIENCE ADANCES | RESEARCH ARTICLE 1 of 11 MATERIALS SCIENCE

Zhang et al., Sci. Adv. 2021; 7 : eabf1966 24 February 2021

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

1 of 11

M A T E R I A L S S C I E N C E

Hierarchical mechanical metamaterials built with scalable tristable elements for ternary logic operation and amplitude modulationHang Zhang1,2, Jun Wu1,2, Daining Fang3*, Yihui Zhang1,2*

Multistable mechanical metamaterials are artificial materials whose microarchitectures offer more than two dif-ferent stable configurations. Existing multistable mechanical metamaterials mainly rely on origami/kirigami- inspired designs, snap-through instability, and microstructured soft mechanisms, with mostly bistable fundamental unit cells. Scalable, tristable structural elements that can be built up to form mechanical metamaterials with an extremely large number of programmable stable configurations remains illusive. Here, we harness the elastic tensile/compressive asymmetry of kirigami microstructures to design a class of scalable X-shaped tristable structures. Using these structure as building block elements, hierarchical mechanical metamaterials with one-dimensional (1D) cylindrical geometries, 2D square lattices, and 3D cubic/octahedral lattices are designed and demonstrated, with capabilities of torsional multistability or independent controlled multidirectional multistability. The number of stable states increases exponentially with the cell number of mechanical metamaterials. The versatile multistability and structural diversity allow demonstrative applications in mechanical ternary logic operators and amplitude modulators with unusual functionalities.

INTRODUCTIONMechanical metamaterials (1–16) represent a type of artificial mate-rials usually consisting of periodic microstructures whose architec-tures are carefully designed to offer mechanical properties that surpass those of conventional materials. These architected metama-terials mainly leverage the spatial motions, extreme deformations, multiple equilibrium states, and shape morphing of microstructures to obtain exotic properties and/or functionalities, such as zero/negative values of Poisson’s ratios (17–20), thermal expansion coefficients (21–24) and swelling ratios (25, 26), reprogrammable stiffness and/or dissipation (27–33), controlled acoustic wave propagation (34, 35), and tailorable multistability (9, 36, 37). The latter property (i.e., tailorable multistability) is of rapidly increasing interests because of promising potentials for applications in information processing (38–41), recyclable energy absorption (42, 43), and soft robotic sys-tems (44–47). These materials, sometimes termed as “multistable mechanical metamaterials,” are designed to offer more than two stable states that can be switched reversibly among each other. This requires an elaborate manipulation of the energy landscape, and several strategies have been reported, including those that rely on origami/kirigami-inspired designs (48–54), snap-through instability (42, 55–59), microstructured soft mechanisms (36, 43, 60), and geo-metrical frustration (61–63). For example, the diversity and high foldability of prismatic geometries have been leveraged in the con-text of origami techniques to develop a class of three-dimensional (3D) multistable metamaterials with periodic arrangements of rigid plates and elastic hinges (49, 52, 53). The instability-based strategies mainly exploited elastic beams capable of snapping between two

different stable configurations to create bistable building block ele-ments that can be further extended to form multistable metamate-rials (40, 42, 55, 64). Despite these important progresses, it remains challenging to design hierarchical metamaterials with thousands of stable states and precisely tailored steady-state properties. In partic-ular, ample opportunities exist in the development of scalable, tristable structural elements that can be built up to form multistable mechanical metamaterials with an extremely high number (e.g., >104) of programmable stable configurations.

This paper introduces a class of X-shaped kirigami microstruc-tures as tristable building block elements, which can be extended, following a bottom-up scheme, to achieve hierarchical mechanical metamaterials with an exponentially increased number of stable states. Here, the tristability arises mainly from the elastic tensile/compressive asymmetry of kirigami microstructures, representing a distinct mechanism from those exploited in previous designs of multistable mechanical metamaterials (36, 37, 40, 42, 43, 55–61). Multimaterial 3D printing technologies enable the fabrication and experimental validation of the programmable multistability in me-chanical metamaterials with 1D cylindrical geometries, 2D square lattices, and 3D cubic/octahedral lattices. The number of stable states in the developed 2D multistable mechanical metamaterials with (M × N) unit cells follows an exponential law of (3M + N) due to the independently controlled multistability along diverse directions. Note that although a few tristable structures have been reported (65–67), they are not scalable and cannot serve as building block elements of multistable mechanical metamaterials. Quantitative studies of the underlying mechanics establish the relationship between the geom-etries of X-shaped kirigami microstructures and the resulting energy landscape. The unique tristable building block structure allows the design and demonstration of fundamental mechanical ternary logic gates, as well as complex logic operators, which are unachievable previously. Compared with binary logic gates, much more informa-tion can be processed through ternary/multivalued logic gates, which can effectively reduce the design complexity and speed up the serial

1AML, Department of Engineering Mechanics, Tsinghua University, Beijing 100084, P.R. China. 2Center for Flexible Electronics Technology, Tsinghua University, Beijing 100084, P.R. China. 3Institute of Advanced Structure Technology, Beijing Key Labo-ratory of Lightweight Multi-Functional Composite Materials and Structures, Beijing Institute of Technology, Beijing 100081, P.R. China.*Corresponding author. Email: [email protected] (D.F.); [email protected] (Y.Z.)

Copyright © 2021 The Authors, some rights reserved; exclusive licensee American Association for the Advancement of Science. No claim to original U.S. Government Works. Distributed under a Creative Commons Attribution NonCommercial License 4.0 (CC BY-NC).

on August 11, 2021

http://advances.sciencemag.org/

Dow

nloaded from

Page 2: MATERIALS SCIENCE Copyright © 2021 Hierarchical mechanical … · Zhang et al., Sci. dv. 2021 7 : eabf1966 24 February 2021 SCIENCE ADANCES | RESEARCH ARTICLE 1 of 11 MATERIALS SCIENCE

Zhang et al., Sci. Adv. 2021; 7 : eabf1966 24 February 2021

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

2 of 11

arithmetic operations (68–70). The ternary logic gates also have unique applications in fuzzy logic circuits (71), asynchronous cir-cuits (72), etc. Furthermore, the elastic tensile/compressive asym-metry of kirigami microstructures and independently controlled tristability of hierarchical metamaterials enable the realization of amplitude modulators capable of filtering low-frequency vibration along different in-plane directions with desired functions, which can be used in noise suppression (73) and nonlinear communication (74). Different from the frequency/phase modulation (75, 76) of vibration, the amplitude modulation is more challenging to realize, because the traditional engineering materials usually show similar mechanical properties under tension and compression.

RESULTSDesign concepts and demonstration of multistable mechanical metamaterials with hierarchical constructionsFigure 1A presents a schematic illustration of the hierarchical design for a multistable mechanical metamaterial consisting of 2D periodic octagonal cells extended in a square array (4 × 4 in this example). Detailed illustrations of the design of the octagonal cell are in fig. S1 and note S1. The octagonal cell can be regarded as a combination of two X-shaped structures, because the X-shaped structure shows the same mechanical responses with that of the structure in the purple frame (Fig. 1A). Such an X-shaped structure serves as a tristable building block structure. Here, the tristability mainly arises from the kirigami microstructures that offer distinct tensile and compressive moduli under uniaxial loadings, such that the X-shaped building block structure can possibly reach an equilibrium state, when the connecting bar (blue) is moved along the x axis. PolyJet multimate-rial 3D printing (fig. S2) allows precise fabrication of the designed multistable mechanical metamaterials (see Materials and Methods for details). Here, the hinges, connecting bars, and substrate are all made of a hard polymer (elastic modulus E ≈ 450 MPa; VeroWhite, Stratasys), while the kirigami microstructures are made of a soft elastomer (elastic modulus E ≈ 0.5 MPa; TangoBlackPlus, Stratasys). Quantitative mechanics modeling of the kirigami microstructures based on finite element analyses (FEAs; see Materials and Methods for details) shows a bending-dominated deformation mechanism under uniaxial stretching (Fig. 1, B and C, and fig. S3), resulting in a much lower tensile modulus (Et) than the compressive modulus (Ec). Experimental measurements of deformed configurations and stress-strain curve show excellent agreement with FEA results, ver-ifying the contrasting tensile/compressive moduli, as evidenced by the modulus ratio of Ec/Et (≈ 101). Owing to such kirigami designs, the X-shaped building block structure can offer two additional stable states with the connecting bar moved leftward/rightward (Fig. 1D). Both the computed and measured load-displacement curves for rightward motion of the connecting bar have an additional mini-mum point (excluding the load-free state), and the corresponding force is negative (i.e., antiparallel with the x axis), pushing the bar further rightward to another equilibrium state (Fig. 1E). The calcu-lated strain energy curve shows three minimum points (I, II, and III), which confirms the tristability of the X-shaped building block structure (Fig. 1E). Note that the modulus ratio (Ec/Et) of the kirigami microstructure should be sufficiently high to ensure the tristability of the X-shaped building block structure (fig. S4).

The tristability of the X-shaped building block structure enables a large number of stable configurations in the hierarchical mechanical

metamaterials. Figure 1F presents all of the nine stable states of the octagonal cell. To decouple the deformations along the x and y directions, we separate the two connecting bars along the out-of-plane direction (i.e., the z direction) in the central region of the octagonal cell (movie S1). This design allows independent control of the stable states along the horizontal and vertical directions through translational motions of the connecting bars (movie S1). Linear array of the octagonal cell along the x and y directions gives rise to a rap-idly increased number of stable states. For example, the mechanical metamaterial with a 2 × 2 array of octagonal cells has 34 (i.e., 81) stable configurations (fig. S5), because the three stable locations of all four connecting bars can be individually addressed. In general, the number of stable states in mechanical metamaterials with M × N array increases exponentially (3M + N) with the total number of con-necting bars. This indicates 6561 (i.e., 38) stable configurations for the mechanical metamaterial with a square array of 4 × 4 octagonal cells. Figure 1G and movie S2 present five different stable states of the fabricated mechanical metamaterial, suggesting the capability of reshaping each octagonal cell into one of the nine possible configu-rations in Fig. 1F. The geometric reconfiguration can be easily implemented through translational motions of connecting bars. Specifically, pushing all the four horizontal bars (along the x axis) rightward in the as-fabricated configuration (① in Fig. 1G) can re-shape the metamaterial into configuration ②. On the basis of con-figuration ②, pushing all the four vertical bars (along the y axis) downward and the bottom horizontal bar back to the initial config-uration leads to the formation of configuration ③. Figure S6 pro-vides 81 stable configurations of this mechanical metamaterial by moving only four connecting bars while fixing the positions of two middle horizontal/vertical bars.

Figure 2 presents the design of multistable mechanical meta-materials with 1D cylindrical geometries and 3D cubic/octahedral lattices. Figure 2A shows the schematic illustration of the cylindrical mechanical metamaterial with torsional multistability, which con-sists of four layers of torsional unit cells. Each torsional unit cell is composed of a driving ring, a constraint ring, a bearing, and an X-shaped tristable building block structure (movie S3). The central region of the X-shaped structure is clamped with the constraint ring and connected to the driving ring through the bearing. This design enables two additional stable configurations by applying clockwise or counterclockwise rotations to the driving ring, as evidenced by the triple-well energy profile (fig. S7). Figure 2B and movie S4 pro-vide experimental and computational results that highlight five different stable configurations of this torsional multistable meta-material. To stabilize at configuration ③ from the initial configura-tion, we rotate the driving rings along the same direction in all of the four torsional unit cells. Because the three stable configurations of each torsional unit cell can be individually controlled, this me-chanical metamaterial offers 34 (i.e., 81) stable configurations in total (fig. S8). In this case, the number of stable states in the me-chanical metamaterials also increases exponentially (3M) with the number (M) of unit cells.

Figure 2C, fig. S9, and note S1 show two representative multi-stable metamaterial designs constructed by extending the octagonal cell (similar to that in Fig. 1) into 3D space, following the cubic and octahedral lattice topologies. Specifically, the architecture of cubic mechanical metamaterial can be obtained by rotating the octagonal cell with respect to the black dashed lines in the leftmost panel (Fig. 2C), followed by connection to the rigid frames (blue color) at

on August 11, 2021

http://advances.sciencemag.org/

Dow

nloaded from

Page 3: MATERIALS SCIENCE Copyright © 2021 Hierarchical mechanical … · Zhang et al., Sci. dv. 2021 7 : eabf1966 24 February 2021 SCIENCE ADANCES | RESEARCH ARTICLE 1 of 11 MATERIALS SCIENCE

Zhang et al., Sci. Adv. 2021; 7 : eabf1966 24 February 2021

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

3 of 11

max-principal

max-principal

d d d d

F[]

EtA U[

]E

tA

y

x

y

z

x

y

x

d

y

z

x

y

x

y

x

y

z

x

A

B C

D

F

G

E

Fig. 1. Design concepts and demonstration of 2D multistable mechanical metamaterials with X-shaped kirigami microstructures. (A) Schematic illustration of the hierarchical construction of a 2D multistable mechanical metamaterial, including the octagonal cells, X-shaped building block structure, and kirigami microstructures. (B) Optical images and FEA results of the kirigami microstructures at undeformed, stretched, and compressed states. (C) Nominal stress-strain curve of the kirigami micro-structure in (B), under both the uniaxial tension and compression. (D) Optical images and FEA results of the three different stable configurations of the 3D-printed X-shaped building block structure. (E) Dependences of the normalized force and the normalized strain energy on the horizontal displacement applied to the X-shaped tristable building block structure in (D). A denotes the cross-sectional area of the microstructure; Ec and Et denote the compressive and tensile moduli, respectively; d denotes the distance marked in (D). (F) Experimental demonstration of the stable configurations of an octagonal cell in the mechanical metamaterial. The red arrows indicate the directions in which the horizontal and vertical connecting bars move. The middle state where no connecting bar moves is marked by a red dashed frame. (G) Experimental demonstration of five representative stable configurations of a 3D-printed mechanical metamaterial with the same geometric parameters as that in (A). Scale bars, 1 mm (B), 5 mm (D and F), and 25 mm (G). Photo credits: Hang Zhang, Tsinghua University.

on August 11, 2021

http://advances.sciencemag.org/

Dow

nloaded from

Page 4: MATERIALS SCIENCE Copyright © 2021 Hierarchical mechanical … · Zhang et al., Sci. dv. 2021 7 : eabf1966 24 February 2021 SCIENCE ADANCES | RESEARCH ARTICLE 1 of 11 MATERIALS SCIENCE

Zhang et al., Sci. Adv. 2021; 7 : eabf1966 24 February 2021

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

4 of 11

the hinges and spatial separation of the connecting bars at the cen-tral region of the cube (see movie S5 for details). Each cubic unit consists of three independently controlled bars, each of which has three stable positions, due to the tristability of the X-shaped build-ing block structure (fig. S10). Thereby, the cubic unit has 33 (i.e., 27) stable configurations (fig. S11). Figure 2D (left) and movie S6 show a few typical stable configurations of a 3D-printed mechanical metamaterial with 2 × 2 × 1 cubic units. Because this metamaterial has eight individually addressable connecting bars, it can offer 38 (i.e., 6561) stable configurations in principle. The octahedral unit (referred to as Case 2 in Fig. 2C) contains six independently con-trolled bars and can thereby render much more stable configura-tions than the cubic unit. Theoretically, the 3D-printed mechanical metamaterial with 2 × 2 × 1 octahedral units (Fig. 2D, right) can offer 320 (i.e., 3,486,784,401) stable configurations, which is inacces-sible previously. The extreme number of stable states holds promise for applications in information processing, as demonstrated by a type of mechanical ternary logic gates and combined logic operators in a subsequent section.

Relationship between mechanical properties and geometrical designs of kirigami microstructures and X-shaped building block structuresUnderstanding of the microstructure-property relationship is essen-tial to the hierarchical design of proposed multistable mechanical metamaterials. Here, we focus on the X-shaped building block struc-ture and establish the connection of its key geometric parameters to the resulting energy landscape. Considering the two-level construc-tion, the geometric parameters can be divided into two categories (one related to the kirigami microstructure and the other to the X-shaped composite). Figure 3A presents a schematic illustration of the kirigami microstructure, where the geometric parameters include the cut lengths, l1 and l2, the cut width c, as well as the overall di-mensions (width a and length b) of the periodical unit. To ensure that the compressive response is very close to that of the parent ma-terial (i.e., to reduce the strain required to result in the self-contact of the microstructures), the cut width should be as small as possible and is fixed as 200 m in this study, considering the precision of exploited commercial 3D printer (layer thickness, ~30 m). The

A

C

1st floor

2nd floor

3rd floor

4th floor

y

x

z

x

y

Case 2

Case 1 Case 1:

Driving ringConstraint ring

BearingHinge

Kirigami microstructure

z

x

yy

x

z

x

yz

x

yz

x

y

Case 2:

z

x

y

x

y

x x

yz y

x

y

x

y

xz

B

D

Connecting bar

Fig. 2. Bottom-up design strategy and demonstration of 3D multistable mechanical metamaterials. (A) Schematic illustration of a torsional multistable mechanical metamaterial consisting of four individually addressable layers. Each layer is composed of a driving ring, a constraining ring, hinges, a bearing, and an X-shaped building block structure. (B) Optical images and FEA results of five representative stable configurations of a 3D-printed torsional mechanical metamaterial with the same geometric parameters as that in (A). (C) Schematic illustration of the cubic and octahedral multistable mechanical metamaterials. The orange and red dashed lines indicate the rota-tion axes of the octagonal cell to form 3D mechanical metamaterials. (D) Experimental demonstration of three representative stable configurations of the 3D-printed cubic and octahedral multistable mechanical metamaterials. Scale bars, 15 mm. Photo credits: Hang Zhang, Tsinghua University.

on August 11, 2021

http://advances.sciencemag.org/

Dow

nloaded from

Page 5: MATERIALS SCIENCE Copyright © 2021 Hierarchical mechanical … · Zhang et al., Sci. dv. 2021 7 : eabf1966 24 February 2021 SCIENCE ADANCES | RESEARCH ARTICLE 1 of 11 MATERIALS SCIENCE

Zhang et al., Sci. Adv. 2021; 7 : eabf1966 24 February 2021

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

5 of 11

dimensionless cut lengths, i.e., ̄ l  1 = l 1 / a and ̄ l  2 = l 2 / a , represent two dominant design parameters that affect the mechanical proper-ties of kirigami microstructures. The other two dimensionless pa-rameters are fixed as ̄ b  = b / a = 0.2 and ̄ c  = c / a = 0.02 in this

set of analyses (Fig. 3), noting that their effects on the stress-strain curve of the kirigami microstructure are illustrated in fig. S12. On the basis of the energy method and incompressible Mooney-Rivlin law, a finite-deformation theoretical model can be developed to

d

L

L

max-principal

max-principalmax-principal

la l

b

c

A

C

F

H

B

D

G

I

E

J

Fig. 3. Microstructure-property relationship of the X-shaped building block structure. (A) Schematic illustration of the kirigami microstructure and the key design param-eters. (B) Experimental and FEA results of the tensile stress-strain curves of the kirigami microstructure with a range of different normalized cut lengths (  ̄ l   1 = l  1 / a and  ̄ l   2 = l  2 / a ). (C) Contour plot of the effective elastic modulus of the kirigami microstructure with respect to the normalized cut lengths (  ̄ l   1 and  ̄ l   2 ). (D) Experimental and FEA results of tensile and compressive stress-strain curves of the kirigami microstructure with homogeneous and composite designs. (E) Optical images and FEA results of the composite kirigami microstructure at different loading states [marked in (D)]. (F) Schematic illustration of the X-shaped building block structure. The key design parameters include the modulus ratio ( = Ec/Et) of the kirigami microstructure under compression to that under tension, the angle of the X-shaped structure, and the length ratio (L/L0). (G) Optical images and FEA results of the two stable configurations of X-shaped building block structures with = 25° and 40° (left and right) for fixed length ratio (L/L0 = 0.64). (H) Load- displacement curves of the homogeneous X-shaped building block structure with different angles (), for fixed modulus ratio ( = 101) and length ratio (L/L0 = 0.64). (I) Similar results in the case of different length ratios (L/L0) for fixed modulus ratio ( = 101) and angle ( = 30°). (J) Load-displacement curves of the composite X-shaped building block structure with different angles () for fixed modulus ratio ( = 240) and length ratio (L/L0 = 0.64). Scale bars, 1 mm (E) and 5 mm (G). Photo credits: Hang Zhang, Tsinghua University.

on August 11, 2021

http://advances.sciencemag.org/

Dow

nloaded from

Page 6: MATERIALS SCIENCE Copyright © 2021 Hierarchical mechanical … · Zhang et al., Sci. dv. 2021 7 : eabf1966 24 February 2021 SCIENCE ADANCES | RESEARCH ARTICLE 1 of 11 MATERIALS SCIENCE

Zhang et al., Sci. Adv. 2021; 7 : eabf1966 24 February 2021

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

6 of 11

predict the stress-strain curve of the kirigami microstructure (fig. S13 and note S2). Figure 3B provides the measured and computed ten-sile stress-strain curves of kirigami microstructures with six differ-ent groups of cut lengths ( ̄ l  1 and ̄ l  2 ). An excellent linear mechanical response can be observed, up to ~80% strain for ̄ l  2 = 0.2 and ~40% strain for ̄ l  2 = 0.4 , which can be partially attributed to the relative linear stress-strain curve of the 3D-printed elastomer material (TangoBlackPlus) (fig. S14). The theoretical and FEA results are in good agreement with the experiments, indicating the mechanics model and FEA as reliable tools to guide the design of kirigami microstructures. The tensile stress-strain curves of kirigami micro-structures with larger cut lengths of ̄ l  2 are shown in fig. S15, where the deviations of FEA and experimental results at large strains are attributed mainly to the fracture at the ends of cuts. Figure 3C shows the contour plot of the initial elastic modulus of kirigami micro-structures in terms of cut lengths ̄ l  1 and ̄ l  2 . The tensile elastic modulus basically increases with decreasing ̄ l  1 or increasing ̄ l  2 because of the increased length of bending-dominated segments. To increase the compressive modulus of kirigami microstructures, the connection region (highlighted in red, Fig. 3D) can be replaced by hard poly-mers (VeroWhitePlus). Figure 3D shows the stress-strain curve of the 3D-printed composite kirigami microstructure ( ̄ l  1 = 0.8 and ̄ l  2 = 0.4 ) under both the uniaxial tension and compression. Compared with the homogeneous design with the same geometric parameters, the compressive elastic modulus of the composite design is increased from 0.5 to 1.2 MPa, while the tensile mechanical responses are fairly close, leading to an increased modulus ratio (Ec/Et) (from 101 to 240). In this case, the simulated microstructure deformations under ten-sion and compression also agree well with the optical images (Fig. 3E).

For a prescribed kirigami microstructure design (with modulus ratio Ec/Et), two additional design parameters govern the energy landscape of the X-shaped building block structure, including the angle of the X-shaped structure and the length ratio (L/L0) of the kirigami microstructure to the total length (Fig. 3F). On the basis of the above energetic model of kirigami microstructures, the princi-ple of virtual work can be used to calculate the reaction force for a prescribed displacement, thereby allowing the prediction of load- displacement curves of the X-shaped building block structure. Figure 3 (H and I) elucidates the effects of these two parameters on mechanical responses of the X-shaped building block structure with a homo-geneous kirigami design (Ec/Et = 101). As the angle increases, the force required to trigger the switch of stable state increases (Fig. 3H), resulting in an increased energy barrier Ebarrier of the landscape (fig. S16A). Note that the effect of friction at hinges is not very evi-dent because of the relatively small area of contact interfaces (<4%, relative to the in-plane projection area of the mechanical metama-terial). For the X-shaped building block structure with the same length ratio (L/L0), the increased angle () also enables a distinct configuration of the higher-order stable state, as evidenced by the larger displacement of the equilibrium state (Fig. 3G). In addition, the X-shaped building block structures can survive high levels of de-formations, as evidenced by the large stretchability (or compressibility) measured experimentally (fig. S17A). The excellent deformability can ensure the reliable switch of the stable state without any failure. Note that the tristability disappears for  < 24° because of the van-ishing energy barrier. Figure 3I shows that the length ratio (L/L0) of the kirigami microstructure only affects the energy barrier Ebarrier and plays a negligible role on the displacement of the higher-order stable states (fig. S16B). A lower length ratio (L/L0) is preferred to

increase the energy barrier, thereby enhancing the stability of higher- order modes. Figure 3J and fig. S16C illustrate that the use of com-posite kirigami design in the X-shaped building block structure yields an enlarged energy barrier and improved tristability, in comparison to that with a homogeneous kirigami design. Specifically, the mod-ulus ratio (i.e., 240) of the composite kirigami microstructure con-taining harder materials is much larger than that (i.e., 101) of the homogeneous design (Fig.  3D). For the X-shaped building block structure with  = 20°, the use of composite kirigami design enables the tristability, while the use of homogeneous kirigami design does not. For the X-shaped building block structures with  = 30° or 35°, the use of composite kirigami design results in a larger force required to switch to high-order stable state, as compared with the case of homogeneous kirigami design (Fig. 3J). The theoretical results show reasonable agreements with the FEA and theoretical results (Fig. 3, H to J). The relatively large derivations near the peak of the load- displacement curve (e.g., in the case of  = 30°, L/L0 = 0.38, and  = 101) are mainly due to the large nominal strain (i.e., >100%) of kirigami microstructure, where the theoretically calculated stress-strain curves of the kirigami microstructure show large discrepancies from the FEA and experimental results. In addition, detailed discussions on the applicability of the FEA methods are provided in note S3.

Mechanical ternary logic gatesThe design flexibility of the X-shaped tristable building block struc-ture facilitates the application in mechanical ternary logic operation, which is unachievable using bistable building block structures (39, 40). For example, in the mechanical system presented by Raney et al. (40), it is very challenging to combine many basic gates for complex logic operations. In comparison to previously reported works (39, 40), the ternary logic gates presented here exploit modular designs, which facilitates the extension of basic gates for complex operations. In addition, compared with the binary logic operation, the ternary log-ic operation can transmit a larger amount of information and ex-ploit a reduced number of basic gates to complete the same logic operation, showing advantages for applications in fuzzy logic (71) and arithmetic and signal processing (68).

Figure 4A presents the schematic illustration of a mechanical ternary NOT gate that is composed of two modules, including an analog-to-digital converter and a digital displacement processor (fig. S18). The analog-to-digital converter is realized by the X-shaped tristable building block structure that converts the continuous dis-placement input into three discrete values. Specifically, the small displacement input that leads to the first-order stable state (initial configuration) of the X-shaped structure is indicated by 0, and the large negative/positive displacement inputs corresponding to two symmetric high-order stable states (deformed configurations) are indicated by −1 and 1, respectively. The digital displacement pro-cessor is realized by the mechanical conversion design proposed by Merkle (77), which can reverse the direction of the input displace-ment. A T-shaped groove is introduced in the substrate to limit the out-of-plane (i.e., z direction) displacements of the converter. Fig-ure 4B and movie S7 provide the experimental demonstration of the operations of the fabricated NOT gate. The ternary logic operations of AND (A ∧ B) and OR (A ∨ B) (Fig. 4C) are more complicated than the binary operations, and the corresponding mechanical gates can be achieved by introducing strategic designs of digital displace-ment processor. Basically, the key is to enable the conversion of two input signals into a desired output signal, according to the truth

on August 11, 2021

http://advances.sciencemag.org/

Dow

nloaded from

Page 7: MATERIALS SCIENCE Copyright © 2021 Hierarchical mechanical … · Zhang et al., Sci. dv. 2021 7 : eabf1966 24 February 2021 SCIENCE ADANCES | RESEARCH ARTICLE 1 of 11 MATERIALS SCIENCE

Zhang et al., Sci. Adv. 2021; 7 : eabf1966 24 February 2021

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

7 of 11

tables (Fig. 4C). Specifically, the output of the AND gate should de-pend on the minimum of two inputs, while the output of the OR gate should depend on the maximum of two inputs. For example, the digital displacement processor of the AND gate is composed of three parts (Fig. 4D and fig. S19): (i) a converter that serves to transform

two displacements into a single output displacement (i.e., at the middle of the converter); (ii) latches that push the converter to move and restrict the converter from being pulled by the kirigami microstructures; and (iii) kirigami microstructures that pull the con-verter back when the two latches both pull the converter. Figure 4D

E –101

–101

D –101

–101

F

–101

–101

A

Input

Substrate

Output

B–1

01

–101

B–1 0 1

A

–1 –1 –1 –10 –1 0 01 –1 0 1

B–1 0 1

A

–1 –1 0 10 0 0 11 1 1 1

C

Hinge

Negative

Positive

G

–101

–101

A B

A B

A B C

A

B

C D

Kirigami microstructure

Fig. 4. Design and experimental demonstration of the ternary mechanical logic gates. (A) Schematic illustration of the NOT gate based on the X-shaped tristable structure. (B) Three operation states of the NOT gate. (C) Truth table of the AND and OR logic operations. (D and E) Six representative operation states of the AND and OR gates. (F) Integration of the three basic logic gates to realize a complex logic operation (A ∨ B) ∧ [NOT(C)], with six representative states provided herein. (G) Six represent-ative states of a complex logic operation, [NOT(A ∨ B)] ∧ (C ∧ D), based on the multistable mechanical metamaterials with a square array of 2 × 2 octagonal cells. Scale bars, 10 mm. Photo credits: Hang Zhang, Tsinghua University.

on August 11, 2021

http://advances.sciencemag.org/

Dow

nloaded from

Page 8: MATERIALS SCIENCE Copyright © 2021 Hierarchical mechanical … · Zhang et al., Sci. dv. 2021 7 : eabf1966 24 February 2021 SCIENCE ADANCES | RESEARCH ARTICLE 1 of 11 MATERIALS SCIENCE

Zhang et al., Sci. Adv. 2021; 7 : eabf1966 24 February 2021

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

8 of 11

and movie S7 show the experimental demonstration of the fabricated AND gate for all possible operations. Following a similar strategy, the ternary OR gate (fig. S20) can be designed, and the digital dis-placement processor is also composed of three parts: (i) a converter with two rectangular cavities; (ii) T-shaped bars that pull the con-verter to move and restrict the converter from being pushed by the springs; and (iii) springs that push the converter back when the two T-shaped bars both push the converter. The logic operations are demonstrated in Fig. 4E. The flexibility of the modular design facil-itates complex logic operations based on the above basic gates. Figure 4F and movie S8 provide an example that realizes the operation, (A ∨ B) ∧ [NOT(C)], for three inputs (A, B, and C). As shown in fig. S21A, the digital displacement processor follows from a combination of digital displacement processors of three basic gates. Another exam-ple of the operation, NOT(A ∧ B), is shown in figs. S21B and S22 and movie S8. The quasi-static time domain diagram of these basic and complex logic operations is shown in fig. S23.

The basic logic gates (e.g., AND gate) allows demonstration of a simple voting device with two different working modes (fig. S24). Specifically, a displacement sensor is used to detect the output dis-placement of the AND gate. After calibration, the sensor could gen-erate an output voltage (~24 V) to turn on the light-emitting diode (LED), if it senses an object within a preset range of distance. For the voting device, each voter has three choices, including “Support,” “Abstain,” and “Oppose,” corresponding to the normalized output “1,” “0,” and “−1,” respectively. In the first voting mode (fig. S24C), the LED light is turned on, only when both voters choose “Support.” By changing the preset range of distance that can be sensed by the displacement sensor, the device can be switched to a different vot-ing mode (fig. S24D), in which the LED light is turned on, as long as one voter chooses “Oppose.”

The large number of stable states enabled by the multistable me-chanical metamaterials allows for complex ternary operations of multiple (e.g., >3) inputs. Figure 4G presents a logic operator based on a mechanical metamaterial (with 2 × 2 array of octagonal cells) that serves as the analog-to-digital converter. Integrated with a spe-cially designed digital displacement processor, this device can real-ize a complex target operation ([NOT(A ∨ B)] ∧ (C ∧ D)) for four different inputs (Fig. 4G and figs. S25A and S26). As shown in fig. S25A, the analog-to-digital converter consists of a 2 × 2 array of octagonal cells. The digital displacement processor of the OR gate is adopted in the x direction, and the digital displacement processor of the NOT gate is used to convert the displacement in the x direction to the y direction. Another two digital displacement processors of the AND gate are used to combine the inputs into a single output. This type of logic operator can also enable the parallel processing of the inputs along different directions, leading to two independent displacement outputs (figs. S25B and S27).

Amplitude modulation of the low-frequency vibrationFigure 5A presents the design of a bidirectional amplitude modulator with developed multistable mechanical metamaterial. Here, the ki-rigami microstructures connected directly to the input ports serve to weaken the transmission of associated forces for displacement loading, while the mechanical metamaterial combines the transmitted force with the tristable units to achieve a regulated displacement output. Figure  5  (B  to  D) shows three representative operational cases of the device for the filtering of low-frequency (i.e., ~0.017 Hz) vibration. For low vibration amplitudes (e.g., 15 mm), the amplitude

modulator filters the triangular wave into a truncated triangular wave (Fig. 5B and movie S9), because the small tensile modulus of the kirigami microstructure almost blocks the transmission of the negative input displacement (Fig. 5B, bottom). As a result, the neg-ative input displacement is suppressed tremendously (attenuation rate, ~95%), while the positive input displacement is transmitted with a relative high fidelity (attenuation rate, ~30%). For medium levels of vibration amplitude (e.g., 25 mm), the modulator filters the tri-angular wave into a step wave (Fig. 5C and movie S9). In this case, the positive input displacement triggers the switch of the octagonal cell into a high-order stable state, while the force associated with the negative input is not large enough to recover the octagonal cell to the original state. Here, we use the critical vibration amplitude (Acr) to denote the displacement at which the switch of stable states oc-curs during vibration. The critical vibration amplitude depends pri-marily on the displacement of the first zero-crossing point of the load-displacement curve that corresponds to the peak point of the energy-displacement curve (see note S4 for details). In the specific example shown in Fig. 5C, the critical vibration amplitude (Acr) to trigger this mode switch of stable state is ~20 mm (note S4 and fig. S28A). For high levels of vibration amplitude (e.g., 55 mm), the large positive input displacement moves the output port along the posi-tive direction further, resulting in sharp conical bulges in the curves of the output displacement (fig. S28B). With a further increase in the vibration amplitude (e.g., to 60 mm), the structure can be pulled from the high-order stable state back to the initial state. On the basis of the above critical vibration amplitude, for an asymmetric vibra-tion with the maximum positive and negative input displacements fixed as 10 and 30 mm, respectively, the triangle wave can be filtered into an output close to a square wave (Fig. 5D and movie S9). The ratios of rise time (~7.0 s) and fall time (~1.5 s) to the vibration peri-od are ~11.7 and 2.5%, respectively, representing a fast mode switch. Here, the amplitude of the square wave depends only on the dis-placement of the equilibrium state of the octagonal cell. The bottom panels (V, VI, and VII) correspond to the deformed configurations marked in the curves. Because of the symmetry of the octagonal cell, the mechanical responses (fig. S29) for the vibration along a vertical direction (i.e., y direction) are very close to that along a horizontal direction (i.e., x direction) in Fig. 5 (B to D), suggesting that the modulator here can work well in both in-plane directions. In addi-tion, the amplitude modulator shows a good durability, as evi-denced by the cyclic testing, where the displacement output is quite stable after 6000 cycles, for the vibration (triangular wave form) with an amplitude of ~15 mm and a frequency of ~0.028 Hz (fig. S30). For vibration with a higher frequency (e.g., 0.05 and 0.1 Hz), the modulator can achieve a similar filtering function (fig. S28, C to E). Here, the overall size (~70 mm) of the octagonal cell and the loading rate (up to 2 mm/s) of the testing equipment based on the stepping motor set a practical limit to the frequency that can be achieved in experiments. Theoretically, the damping of the kirigami microstructure could also affect the frequency limit of the system, because the 3D-printed rubber material (i.e., TangoBlackPlus) used in the current work shows a certain level of viscoelastic behavior at room temperature. This could be addressed by using a lower- damping material (with a reduced viscoelasticity) in the fabrication. Because the developed amplitude modulator mainly leverages the multi-stability of the octagonal cell and the different tensile/compressive stiff-ness of the kirigami microstructure, similar filtering function can be expected for inputs with other wave forms (e.g., sinusoidal waves).

on August 11, 2021

http://advances.sciencemag.org/

Dow

nloaded from

Page 9: MATERIALS SCIENCE Copyright © 2021 Hierarchical mechanical … · Zhang et al., Sci. dv. 2021 7 : eabf1966 24 February 2021 SCIENCE ADANCES | RESEARCH ARTICLE 1 of 11 MATERIALS SCIENCE

Zhang et al., Sci. Adv. 2021; 7 : eabf1966 24 February 2021

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

9 of 11

Typically, for robots in harsh environment (e.g., high radiation and strong magnetic fields), electronic devices would not work properly, and mechanical devices could be a good alternative. The amplitude modulator can be used under these circumstances to filter different input signals into desired output signals. For example, the modula-tor could serve as a displacement detector to sense whether the in-put displacement exceeds a given threshold or a damping platform that reduces susbtantially the influence of the input displacement on the output. The modulation of vibration amplitude demonstrat-ed here can be potentially used in noise suppression (73) and non-linear communication (74).

DISCUSSIONThis paper reports the design, fabrication, and characterization of a class of hierarchical mechanical metamaterials with an expo-nentially increased number of stable states. Starting from the pro-grammable X-shaped tristable building block structure, hierarchical

mechanical metamaterials with 1D cylindrical geometries, 2D square lattices, and 3D cubic/octahedral lattices are designed and demon-strated, with capabilities of torsional multistability or indepen-dent controlled multidirectional multistability. Validated mechanics modeling serves as the basis of the design strategy and sheds light on the underlying relationship between the microstructural ge-ometries and the resulting energy landscape. The design flexibil-ity and ample stable states allow demonstrative applications in mechanical ternary logic gates, including three basic gates (i.e., NOT, AND, and OR gates) and their combined logic operations. We also harness the structural diversity and independently con-trolled tristability to realize a type of amplitude modulator that can filter low-frequency vibration of different amplitudes into different wave forms (e.g., truncated triangular wave, step wave, and square wave). These mechanical devices hold promising potentials for uses in the motion/configuration control of soft actuators and ro-botics. Compared with traditional electrical devices, these mechan-ical devices show advantages in energy saving (77) and resistance to

x

x

x

x

x

x

x

y

y

x

z

x

y

A

B C D

Fig. 5. Applications in the amplitude modulation of the low-frequency vibration. (A) Conceptual illustration of the modular design of the amplitude modulator. Here, the amplitude modulator works along the x and y axes, and the sign of input/output displacements is consistent with the sign of the coordinate axes. The module 1 serves to weaken the force transmission, and the module 2 combines the transmitted force with the tristable units to achieve a regulated displacement output. (B) Input and output displacements along the x direction versus the time for a low level of amplitude, showing the function of filtering the triangular wave as the truncated triangular wave. The optical images on the bottom panel correspond to the two states marked in the curves. (C) Similar results in the case of an intermediate level of amplitude, showing the function of filtering the triangular wave as the step wave. (D) Similar results in the case of a high level of amplitude, showing the function of filtering the triangular wave as the square wave. Scale bars, 15 mm. Photo credits: Hang Zhang, Tsinghua University.

on August 11, 2021

http://advances.sciencemag.org/

Dow

nloaded from

Page 10: MATERIALS SCIENCE Copyright © 2021 Hierarchical mechanical … · Zhang et al., Sci. dv. 2021 7 : eabf1966 24 February 2021 SCIENCE ADANCES | RESEARCH ARTICLE 1 of 11 MATERIALS SCIENCE

Zhang et al., Sci. Adv. 2021; 7 : eabf1966 24 February 2021

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

10 of 11

corrosion (78) in a harsh environment (e.g., high radiation and strong magnetic fields).

MATERIALS AND METHODSFabrication and mechanical testing of multistable mechanical metamaterialsAll the samples in this study were prepared by multimaterials PolyJet 3D printing (Objet350, Stratasys). Two different materials (VeroWhitePlus and TangoBlackPlus) available through multimaterial 3D printers (Objet350, Stratasys) were adopted, and both were cured by ultra-violet lamps at room temperature. TangoBlackPlus is a soft elastomer (~0.5 MPa), mainly composed of exo-1,7,7-trimethylbicyclo[2.2.1]hept-2-yl acrylate and photoinitiators. VeroWhitePlus is a harder polymer (~680 MPa), mainly composed of exo-1,7,7-trimethylbicyclo[2.2.1]hept- 2-yl acrylate, tricyclodecane dimethanol diacrylate, titanium dioxide, and photoinitiators. A digital camera (760D, Canon) recorded the deformation process of the mechanical metamaterials, as well as the operations of the ternary logic operators and the amplitude modu-lator. The load-displacement curves of the steady-state unit were obtained by a commercial mechanical testing machine. A custom-erized machine enabled the loading and unloading of the amplitude modulator with a constant loading rate of preset magnitude.

Finite element analysesThe commercial software ABAQUS (SIMULIA) was used to perform the FEAs. We adopted an implicit solver to calculate the deformations of the multistable mechanical metamaterials, the tensile/compressive stress-strain curves of the kirigami microstructures, and the load- displacement curves of X-shaped building block structure. The geo-metric nonlinearity was taken into account in the calculations. The FEA exploited eight-node linear hybridization elements, with refined meshes to ensure the convergence of the calculations. Because the elastic modulus of VeroWhitePlus is much larger than that of TangoBlackPlus, the strain in VeroWhitePlus during deformation is typically very small (<0.1%). Therefore, a linear elastic constitu-tive relationship (EVeroWhitePlus = 680 MPa, vVeroWhitePlus = 0.4) was used to simplify the simulation. Differently, the incompressible Mooney- Rivlin hyperelastic constitutive relationship was used to model the deformations of the elastomer (TangoBlackPlus), where the uniaxial stress-strain curve measured by the experiment was imported into the simulations.

SUPPLEMENTARY MATERIALSSupplementary material for this article is available at http://advances.sciencemag.org/cgi/content/full/7/9/eabf1966/DC1

REFERENCES AND NOTES 1. A. A. Zadpoor, Mechanical meta-materials. Mater. Horiz. 3, 371–381 (2016). 2. J. R. Greer, V. S. Deshpande, Three-dimensional architected materials and structures:

Design, fabrication, and mechanical behavior. MRS Bull. 44, 750–757 (2019). 3. C. Coulais, D. Sounas, A. Alù, Static non-reciprocity in mechanical metamaterials. Nature

542, 461–464 (2017). 4. X. Li, H. Gao, Smaller and stronger. Nat. Mater. 15, 373–374 (2016). 5. X. Zhang, L. Zhong, A. Mateos, A. Kudo, A. Vyatskikh, H. Gao, J. R. Greer, X. Li, Theoretical

strength and rubber-like behaviour in micro-sized pyrolytic carbon. Nat. Nanotechnol. 14, 762–769 (2019).

6. K. Bertoldi, V. Vitelli, J. Christensen, M. van Hecke, Flexible mechanical metamaterials. Nat. Rev. Mater. 2, 17066 (2017).

7. Y. Zhang, F. Zhang, Z. Yan, Q. Ma, X. Li, Y. Huang, J. A. Rogers, Printing, folding and assembly methods for forming 3D mesostructures in advanced materials. Nat. Rev. Mater. 2, 17019 (2017).

8. L. R. Meza, A. J. Zelhofer, N. Clarke, A. J. Mateos, D. M. Kochmann, J. R. Greer, Resilient 3D hierarchical architected metamaterials. Proc. Natl. Acad. Sci. U.S.A. 112, 11502–11507 (2015).

9. D. M. Kochmann, J. B. Hopkins, L. Valdevit, Multiscale modeling and optimization of the mechanics of hierarchical metamaterials. MRS Bull. 44, 773–781 (2019).

10. R. Gatt, L. Mizzi, J. I. Azzopardi, K. M. Azzopardi, D. Attard, A. Casha, J. Briffa, J. N. Grima, Hierarchical auxetic mechanical metamaterials. Sci. Rep. 5, 8395 (2015).

11. D. Mousanezhad, S. Babaee, H. Ebrahimi, R. Ghosh, A. S. Hamouda, K. Bertoldi, A. Vaziri, Hierarchical honeycomb auxetic metamaterials. Sci. Rep. 5, 18306 (2015).

12. W. Yang, Q. Liu, Z. Gao, Z. Yue, B. Xu, Theoretical search for heterogeneously architected 2D structures. Proc. Natl. Acad. Sci. U.S.A. 115, E7245–E7254 (2018).

13. Y. Tang, G. Lin, L. Han, S. Qiu, S. Yang, J. Yin, Design of hierarchically cut hinges for highly stretchable and reconfigurable metamaterials with enhanced strength. Adv. Mater. 27, 7181–7190 (2015).

14. Y. Tang, G. Lin, S. Yang, Y. K. Yi, R. D. Kamien, J. Yin, Programmable kiri-kirigami metamaterials. Adv. Mater. 29, 1604262 (2017).

15. S. Lim, D. Son, J. Kim, Y. B. Lee, J.-K. Song, S. Choi, D. J. Lee, J. H. Kim, M. Lee, T. Hyeon, D.-H. Kim, Transparent and stretchable interactive human machine interface based on patterned graphene heterostructures. Adv. Funct. Mater. 25, 375–383 (2015).

16. E. T. Filipov, T. Tachi, G. H. Paulino, Origami tubes assembled into stiff, yet reconfigurable structures and metamaterials. Proc. Natl. Acad. Sci. U.S.A. 112, 12321–12326 (2015).

17. M. Lei, W. Hong, Z. Zhao, C. Hamel, M. Chen, H. Lu, H. J. Qi, 3D printing of auxetic metamaterials with digitally reprogrammable shape. ACS Appl. Mater. Interfaces 11, 22768–22776 (2019).

18. Y. Jiang, Z. Liu, N. Matsuhisa, D. Qi, W. R. Leow, H. Yang, J. Yu, G. Chen, Y. Liu, C. Wan, Z. Liu, X. Chen, Auxetic mechanical metamaterials to enhance sensitivity of stretchable strain sensors. Adv. Mater. 30, 1706589 (2018).

19. Z. Zhao, C. Yuan, M. Lei, L. Yang, Q. Zhang, H. Chen, H. J. Qi, D. Fang, Three-dimensionally printed mechanical metamaterials with thermally tunable auxetic behavior. Phys. Rev. Appl. 11, 044074 (2019).

20. P. P. Pratapa, K. Liu, G. H. Paulino, Geometric mechanics of origami patterns exhibiting poisson's ratio switch by breaking mountain and valley assignment. Phys. Rev. Lett. 122, 155501 (2019).

21. X. Ni, X. Guo, J. Li, Y. Huang, Y. Zhang, J. A. Rogers, 2D mechanical metamaterials with widely tunable unusual modes of thermal expansion. Adv. Mater. 31, 1905405 (2019).

22. R. Lakes, Cellular solids with tunable positive or negative thermal expansion of unbounded magnitude. Appl. Phys. Lett. 90, 221905 (2007).

23. J. N. Grima, B. Ellul, D. Attard, R. Gatt, M. Attard, Composites with needle-like inclusions exhibiting negative thermal expansion: A preliminary investigation. Compos. Sci. Technol. 70, 2248–2252 (2010).

24. E. Boatti, N. Vasios, K. Bertoldi, Origami metamaterials for tunable thermal expansion. Adv. Mater. 29, 1700360 (2017).

25. J. Liu, T. Gu, S. Shan, S. H. Kang, J. C. Weaver, K. Bertoldi, Harnessing buckling to design architected materials that exhibit effective negative swelling. Adv. Mater. 28, 6619–6624 (2016).

26. H. Zhang, X. Guo, J. Wu, D. Fang, Y. Zhang, Soft mechanical metamaterials with unusual swelling behavior and tunable stress-strain curves. Sci. Adv. 4, eaar8535 (2018).

27. M. J. Mirzaali, H. Pahlavani, A. A. Zadpoor, Auxeticity and stiffness of random networks: Lessons for the rational design of 3D printed mechanical metamaterials. Appl. Phys. Lett. 115, 021901 (2019).

28. R. Khajehtourian, D. M. Kochmann, Phase transformations in substrate-free dissipative multistable metamaterials. Extreme Mech. Lett. 37, 100700 (2020).

29. S. N. Khaderi, V. S. Deshpande, N. A. Fleck, The stiffness and strength of the gyroid lattice. Int. J. Solids Struct. 51, 3866–3877 (2014).

30. D. Yan, J. Chang, H. Zhang, J. Liu, H. Song, Z. Xue, F. Zhang, Y. Zhang, Soft three-dimensional network materials with rational bio-mimetic designs. Nat. Commun. 11, 1180 (2020).

31. Z. Zhai, Y. Wang, H. Jiang, Origami-inspired, on-demand deployable and collapsible mechanical metamaterials with tunable stiffness. Proc. Natl. Acad. Sci. U.S.A. 115, 2032–2037 (2018).

32. C. Cao, X. Zhao, Tunable stiffness of electrorheological elastomers by designing mesostructures. Appl. Phys. Lett. 103, 041901 (2013).

33. K. Liu, T. Zegard, P. P. Pratapa, G. H. Paulino, Unraveling tensegrity tessellations for metamaterials with tunable stiffness and bandgaps. J. Mech. Phys. Solids 131, 147–166 (2019).

34. A. Palermo, Y. Wang, P. Celli, C. Daraio, Tuning of surface-acoustic-wave dispersion via magnetically modulated contact resonances. Phys. Rev. Appl. 11, 044057 (2019).

35. Y. Wang, B. Yousefzadeh, H. Chen, H. Nassar, G. Huang, C. Daraio, Observation of nonreciprocal wave propagation in a dynamic phononic lattice. Phys. Rev. Lett. 121, 194301 (2018).

on August 11, 2021

http://advances.sciencemag.org/

Dow

nloaded from

Page 11: MATERIALS SCIENCE Copyright © 2021 Hierarchical mechanical … · Zhang et al., Sci. dv. 2021 7 : eabf1966 24 February 2021 SCIENCE ADANCES | RESEARCH ARTICLE 1 of 11 MATERIALS SCIENCE

Zhang et al., Sci. Adv. 2021; 7 : eabf1966 24 February 2021

S C I E N C E A D V A N C E S | R E S E A R C H A R T I C L E

11 of 11

36. B. Haghpanah, L. Salari-Sharif, P. Pourrajab, J. Hopkins, L. Valdevit, Multistable shape-reconfigurable architected materials. Adv. Mater. 28, 7915–7920 (2016).

37. L. Jin, R. Khajehtourian, J. Mueller, A. Rafsanjani, V. Tournat, K. Bertoldi, D. M. Kochmann, Guided transition waves in multistable mechanical metamaterials. Proc. Natl. Acad. Sci. U.S.A. 117, 2319–2325 (2020).

38. R. C. Merkle, R. A. Freitas Jr., T. Hogg, T. E. Moore, M. S. Moses, J. Ryley, Mechanical computing systems using only links and rotary joints. J. Mech. Robot. 10, 061006 (2018).

39. Y. Song, R. M. Panas, S. Chizari, L. A. Shaw, J. A. Jackson, J. B. Hopkins, A. J. Pascall, Additively manufacturable micro-mechanical logic gates. Nat. Commun. 10, 882 (2019).

40. J. R. Raney, N. Nadkarni, C. Daraio, D. M. Kochmann, J. A. Lewis, K. Bertoldi, Stable propagation of mechanical signals in soft media using stored elastic energy. Proc. Natl. Acad. Sci. U.S.A. 113, 9722–9727 (2016).

41. D. J. Preston, P. Rothemund, H. J. Jiang, M. P. Nemitz, J. Rawson, Z. Suo, G. M. Whitesides, Digital logic for soft devices. Proc. Natl. Acad. Sci. U.S.A. 116, 7750–7759 (2019).

42. T. Frenzel, C. Findeisen, M. Kadic, P. Gumbsch, M. Wegener, Tailored buckling microlattices as reusable light-weight shock absorbers. Adv. Mater. 28, 5865–5870 (2016).

43. S. Shan, S. H. Kang, J. R. Raney, P. Wang, L. Fang, F. Candido, J. A. Lewis, K. Bertoldi, Multistable architected materials for trapping elastic strain energy. Adv. Mater. 27, 4296–4301 (2015).

44. G. Gu, J. Zou, R. Zhao, X. Zhao, X. Zhu, Soft wall-climbing robots. Sci. Robot. 3, eaat2874 (2018).

45. Y. Kim, G. A. Parada, S. Liu, X. Zhao, Ferromagnetic soft continuum robots. Sci. Robot. 4, eaax7329 (2019).

46. N. Lu, D.-H. Kim, Flexible and stretchable electronics paving the way for soft robotics. Soft Robot. 1, 53–62 (2013).

47. P. Bhovad, J. Kaufmann, S. Li, Peristaltic locomotion without digital controllers: Exploiting multi-stability in origami to coordinate robotic motion. Extreme Mech. Lett. 32, 100552 (2019).

48. F. S. L. Bobbert, S. Janbaz, T. van Manen, Y. Li, A. A. Zadpoor, Russian doll deployable meta-implants: Fusion of kirigami, origami, and multi-stability. Mater. Des. 191, 108624 (2020).

49. J. T. B. Overvelde, J. C. Weaver, C. Hoberman, K. Bertoldi, Rational design of reconfigurable prismatic architected materials. Nature 541, 347–352 (2017).

50. G. P. T. Choi, L. H. Dudte, L. Mahadevan, Programming shape using kirigami tessellations. Nat. Mater. 18, 999–1004 (2019).

51. L. H. Dudte, E. Vouga, T. Tachi, L. Mahadevan, Programming curvature using origami tessellations. Nat. Mater. 15, 583–588 (2016).

52. J. L. Silverberg, J.-H. Na, A. A. Evans, B. Liu, T. C. Hull, C. D. Santangelo, R. J. Lang, R. C. Hayward, I. Cohen, Origami structures with a critical transition to bistability arising from hidden degrees of freedom. Nat. Mater. 14, 389–393 (2015).

53. J. L. Silverberg, A. A. Evans, L. McLeod, R. C. Hayward, T. Hull, C. D. Santangelo, I. Cohen, Using origami design principles to fold reprogrammable mechanical metamaterials. Science 345, 647–650 (2014).

54. K. Liu, T. Tachi, G. H. Paulino, Invariant and smooth limit of discrete geometry folded from bistable origami leading to multistable metasurfaces. Nat. Commun. 10, 4238 (2019).

55. A. Rafsanjani, A. Akbarzadeh, D. Pasini, Snapping mechanical metamaterials under tension. Adv. Mater. 27, 5931–5935 (2015).

56. P. M. Reis, A perspective on the revival of structural (in)stability with novel opportunities for function: From buckliphobia to buckliphilia. J. Appl. Mech. 82, 111001 (2015).

57. H. Fu, K. Nan, W. Bai, W. Huang, K. Bai, L. Lu, C. Zhou, Y. Liu, F. Liu, J. Wang, M. Han, Z. Yan, H. Luan, Y. Zhang, Y. Zhang, J. Zhao, X. Cheng, M. Li, J. W. Lee, Y. Liu, D. Fang, X. Li, Y. Huang, Y. Zhang, J. A. Rogers, Morphable 3D mesostructures and microelectronic devices by multistable buckling mechanics. Nat. Mater. 17, 268–276 (2018).

58. B. Florijn, C. Coulais, M. van Hecke, Programmable mechanical metamaterials. Phys. Rev. Lett. 113, 175503 (2014).

59. J. Marthelot, P.-T. Brun, F. L. Jiménez, P. M. Reis, Reversible patterning of spherical shells through constrained buckling. Phys. Rev. Mater. 1, 025601 (2017).

60. F. Pan, Y. Li, Z. Li, J. Yang, B. Liu, Y. Chen, 3D pixel mechanical metamaterials. Adv. Mater. 31, 1900548 (2019).

61. C. Coulais, E. Teomy, K. de Reus, Y. Shokef, M. van Hecke, Combinatorial design of textured mechanical metamaterials. Nature 535, 529–532 (2016).

62. S. H. Kang, S. Shan, A. Košmrlj, W. L. Noorduin, S. Shian, J. C. Weaver, D. R. Clarke, K. Bertoldi, Complex ordered patterns in mechanical instability induced geometrically frustrated triangular cellular structures. Phys. Rev. Lett. 112, 098701 (2014).

63. S. Waitukaitis, R. Menaut, B. G.-g. Chen, M. van Hecke, Origami multistability: From single vertices to metasheets. Phys. Rev. Lett. 114, 055503 (2015).

64. H. Yang, L. Ma, 1D to 3D multi-stable architected materials with zero Poisson's ratio and controllable thermal expansion. Mater. Des. 188, 108430 (2020).

65. L.-C. Wang, W.-L. Song, Y.-J. Zhang, M.-J. Qu, Z. Zhao, M. Chen, Y. Yang, H. Chen, D. Fang, Active reconfigurable tristable square-twist origami. Adv. Funct. Mater. 30, 1909087 (2020).

66. X.-Q. Li, W.-B. Li, W.-M. Zhang, H.-X. Zou, Z.-K. Peng, G. Meng, Magnetic force induced tristability for dielectric elastomer actuators. Smart Mater. Struct. 26, 105007 (2017).

67. B. H. Coburn, A. Pirrera, P. M. Weaver, S. Vidoli, Tristability of an orthotropic doubly curved shell. Compos. Struct. 96, 446–454 (2013).

68. Smith, The prospects for multivalued logic: A technology and applications view. IEEE Trans. Comput. C-30, 619–634 (1981).

69. S. Lin, Y. Kim, F. Lombardi, CNTFET-based design of ternary logic gates and arithmetic circuits. IEEE Trans. Nanotechnol. 10, 217–225 (2011).

70. T. Chattopadhyay, All-optical symmetric ternary logic gate. Opt. Laser Technol. 42, 1014–1021 (2010).

71. T. Araki, H. Tatsumi, M. Mukaidono, F. Yamamoto, Minimization of incompletely specified regular ternary logic functions and its application to fuzzy switching functions, Proceedings. 1998 28th IEEE International Symposium on Multiple-Valued Logic (Cat. No.98CB36138) (Fukuoka, Japan, 1998), pp. 289–296.

72. R. Mariani, R. Roncella, R. Saletti, P. Terreni, A useful application of CMOS ternary logic to the realisation of asynchronous circuits. Proc. IEEE Int. Symp. Multiple-Valued Logic, 203–208 (1997).

73. J. Tchorz, B. Kollmeier, SNR estimation based on amplitude modulation analysis with applications to noise suppression. IEEE Trans. Speech Audio Process. 11, 184–192 (2003).

74. N. A. Shevchenko, S. A. Derevyanko, J. E. Prilepsky, A. Alvarado, P. Bayvel, S. K. Turitsyn, Capacity lower bounds of the noncentral chi-channel with applications to soliton amplitude modulation. IEEE Trans. Commun. 66, 2978–2993 (2018).

75. B. Deng, P. Wang, V. Tournat, K. Bertoldi, Nonlinear transition waves in free-standing bistable chains. J. Mech. Phys. Solids 136, 103661 (2020).

76. M. Kafesaki, M. M. Sigalas, N. García, Frequency modulation in the transmittivity of wave guides in elastic-wave band-gap materials. Phys. Rev. Lett. 85, 4044–4047 (2000).

77. R. C. Merkle, Two types of mechanical reversible logic. Nanotechnology 4, 114–131 (1993). 78. A. De Greef, P. Lambert, A. Delchambre, Towards flexible medical instruments: Review

of flexible fluidic actuators. Precis. Eng. 33, 311–321 (2009).

Acknowledgments Funding: Y.Z. acknowledges support from the National Natural Science Foundation of China (grant nos. 11722217 and 11921002), the Tsinghua National Laboratory for Information Science and Technology, the Tsinghua University Initiative Scientific Research Program (grant no. 2019Z08QCX10), and a grant from the Institute for Guo Qiang, Tsinghua University (grant no. 2019GQG1012). Author contributions: Y.Z. designed and supervised the research. Y.Z. and H.Z. led the structural designs, mechanics modeling, and experimental work, with assistance from J.W. Y.Z., H.Z., and D.F. wrote the manuscript and designed the figures. All authors commented on the paper. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials. Additional data related to this paper may be requested from the corresponding authors.

Submitted 13 October 2020Accepted 11 January 2021Published 24 February 202110.1126/sciadv.abf1966

Citation: H. Zhang, J. Wu, D. Fang, Y. Zhang, Hierarchical mechanical metamaterials built with scalable tristable elements for ternary logic operation and amplitude modulation. Sci. Adv. 7, eabf1966 (2021).

on August 11, 2021

http://advances.sciencemag.org/

Dow

nloaded from

Page 12: MATERIALS SCIENCE Copyright © 2021 Hierarchical mechanical … · Zhang et al., Sci. dv. 2021 7 : eabf1966 24 February 2021 SCIENCE ADANCES | RESEARCH ARTICLE 1 of 11 MATERIALS SCIENCE

operation and amplitude modulationHierarchical mechanical metamaterials built with scalable tristable elements for ternary logic

Hang Zhang, Jun Wu, Daining Fang and Yihui Zhang

DOI: 10.1126/sciadv.abf1966 (9), eabf1966.7Sci Adv 

ARTICLE TOOLS http://advances.sciencemag.org/content/7/9/eabf1966

MATERIALSSUPPLEMENTARY http://advances.sciencemag.org/content/suppl/2021/02/22/7.9.eabf1966.DC1

REFERENCES

http://advances.sciencemag.org/content/7/9/eabf1966#BIBLThis article cites 76 articles, 9 of which you can access for free

PERMISSIONS http://www.sciencemag.org/help/reprints-and-permissions

Terms of ServiceUse of this article is subject to the

is a registered trademark of AAAS.Science AdvancesYork Avenue NW, Washington, DC 20005. The title (ISSN 2375-2548) is published by the American Association for the Advancement of Science, 1200 NewScience Advances

License 4.0 (CC BY-NC).Science. No claim to original U.S. Government Works. Distributed under a Creative Commons Attribution NonCommercial Copyright © 2021 The Authors, some rights reserved; exclusive licensee American Association for the Advancement of

on August 11, 2021

http://advances.sciencemag.org/

Dow

nloaded from