Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-42gr6 Total loading time: 0 Render date: 2024-04-18T05:05:46.060Z Has data issue: false hasContentIssue false

15 - Functionalized poly(ethylene glycol) hydrogels for controlling stem cell fate

from Part III - Hydrogel scaffolds for regenerative medicine

Published online by Cambridge University Press:  05 February 2015

Chien-Chi Lin
Affiliation:
Indiana University-Purdue University Indianapolis
Peter X. Ma
Affiliation:
University of Michigan, Ann Arbor
Get access

Summary

Introduction

An emerging trend in tissue engineering and regenerative medicine is the design of scaffolds that recapitulate a stem cell’s extracellular microenvironment, or “niche” (Cushing and Anseth 2007; Tibbitt and Anseth 2009). While the exact constituents of a stem cell niche are not well characterized, it is generally believed that morphogens, immobilized extracellular matrix (ECM) components, cell–cell interactions, and matrix elasticity are the major determinants that direct stem cell self-renewal or differentiation (Lutolf and Hubbell 2005). Research over the years has accumulated fundamental knowledge about stem cell biology in vitro. Most of the results were obtained from experiments conducting on flat, rigid, and two-dimensional (2D) tissue culture polystyrene (TCPS). These 2D environments, however, might not truly reproduce what a stem cell experiences in vivo, where relevant biological and biophysical cues are produced in a far more complex and spatially and temporally regulated manner. In fact, results from numerous studies have demonstrated that, compared with cell culture on TCPS, stem cells behave differently when cultured in three-dimensional (3D) matrices (Liu and Roy 2005; Ingber et al. 2006; Lee et al. 2008b; Lund et al. 2009).

Artificial matrices of synthetic or natural origins have been developed for culturing stem cells in 3D (Lee et al. 2008b; Lund et al. 2009). While natural ECM components, such as collagen and laminin, provide intrinsic biological recognition sites for cell attachment and migration, these natural matrices are ill-defined and poorly controlled with respect to their compositions and mechanical properties, respectively. Furthermore, one cannot afford to overlook the immunogenicity problems associated with allogenic or xenogenic materials, since this significantly limits the clinical translatability/relevance of stem-cell-based therapy. Matrices fabricated from synthetic polymers, on the other hand, provide well-defined and user-controllable chemical and physical properties (Nuttelman et al. 2008). The use of synthetic matrices, however, is not without challenges, insofar as synthetic polymers usually lack biological motifs for cellular recognition. Hence, significant efforts have been dedicated to fabricating synthetic polymeric scaffolds functionalized with biomimetic motifs that permit cellular interaction.

Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2014

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Almany, L. and Seliktar, D. 2005. Biosynthetic hydrogel scaffolds made from fibrinogen and polyethylene glycol for 3D cell cultures. Biomaterials, 26(15), 2467–77.CrossRefGoogle ScholarPubMed
Banerjee, A., Arha, M., Choudhary, S. et al. (2009. The influence of hydrogel modulus on the proliferation and differentiation of encapsulated neural stem cells. Biomaterials, 30(27), 4695–99.CrossRefGoogle ScholarPubMed
Baskin, J. M., Prescher, J. A., Laughlin, S. T. et al. 2007. Copper-free click chemistry for dynamic in vivo imaging. Proc. Nat. Acad. Sci. USA, 104(43), 16793–7.CrossRefGoogle ScholarPubMed
Benoit, D. S. W., Durney, A. R. and Anseth, K. S. 2007. The effect of heparin-functionalized PEG hydrogels on three-dimensional human mesenchymal stem cell osteogenic differentiation. Biomaterials, 28(1), 66–77.CrossRefGoogle ScholarPubMed
Benoit, D. S. W., Nuttelman, C. R., Collins, S. D. and Anseth, K. S. 2006. Synthesis and characterization of a fluvastatin-releasing hydrogel delivery system to modulate hMSC differentiation and function for bone regeneration. Biomaterials, 27(36), 6102–10.CrossRefGoogle ScholarPubMed
Cushing, M. C. and Anseth, K. S. 2007. Hydrogel cell cultures. Science, 316(5828), 1133–4.CrossRefGoogle ScholarPubMed
Cushing, M. C., Liao, J. T., Jaeggli, M. P. and Anseth, K. S. 2007. Material-based regulation of the myofibroblast phenotype. Biomaterials, 28(23), 3378–87.CrossRefGoogle ScholarPubMed
DeForest, C. A., Polizzotti, B. D. and Anseth, K. S. 2009. Sequential click reactions for synthesizing and patterning three-dimensional cell microenvironments. Nature Mater., 8(8), 659–64.CrossRefGoogle ScholarPubMed
DeForest, C. A., Sims, E. A. and Anseth, K. S. 2010. Peptide-functionalized click hydrogels with independently tunable mechanics and chemical functionality for 3D cell culture. Chem. Mater., 22(16), 4783–90.CrossRefGoogle ScholarPubMed
DeLong, S. A., Moon, J. J. and West, J. L. 2005. Covalently immobilized gradients of bFGF on hydrogel scaffolds for directed cell migration. Biomaterials, 26(16), 3227–34.CrossRefGoogle ScholarPubMed
Dikovsky, D., Bianco-Peled, H. and Seliktar, D. 2006. The effect of structural alterations of PEG–fibrinogen hydrogel scaffolds on 3-D cellular morphology and cellular migration. Biomaterials, 27(8), 1496–506.CrossRefGoogle ScholarPubMed
Fairbanks, B. D., Schwartz, M. P., Halevi, A. E. et al. 2009. A versatile synthetic extracellular matrix mimic via thiol–norbornene photopolymerization. Adv. Mater., 21(48), 5005–10.CrossRefGoogle ScholarPubMed
Freudenberg, U., Hermann, A., Welzel, P. B. et al. 2009. A star–PEG–heparin hydrogel platform to aid cell replacement therapies for neurodegenerative diseases. Biomaterials, 30(28), 5049–60.CrossRefGoogle ScholarPubMed
Gonen-Wadmany, M., Goldshmid, R. and Seliktar, D. 2011. Biological and mechanical implications of PEGylating proteins into hydrogel biomaterials. Biomaterials, 32(26), 6025–33.CrossRefGoogle ScholarPubMed
Gonen-Wadmany, M., Oss-Ronen, L. and Seliktar, D. 2007. Protein–polymer conjugates for forming photopolymerizable biomimetic hydrogels for tissue engineering. Biomaterials, 28(26), 3876–86.CrossRefGoogle ScholarPubMed
He, X. Z., Ma, J. Y. and Jabbari, E. 2008. Effect of grafting RGD and BMP-2 protein-derived peptides to a hydrogel substrate on osteogenic differentiation of marrow stromal cells. Langmuir, 24(21), 12508–16.CrossRefGoogle ScholarPubMed
Hern, D. L. and Hubbell, J. A. 1998. Incorporation of adhesion peptides into nonadhesive hydrogels useful for tissue resurfacing. J. Biomed. Mater. Res., 39(2), 266–76.3.0.CO;2-B>CrossRefGoogle ScholarPubMed
Hu, B. H., Su, J. and Messersmith, P. B. 2009. Hydrogels cross-linked by native chemical ligation. Biomacromolecules, 10(8), 2194–200.CrossRefGoogle ScholarPubMed
Hwang, N. S., Varghese, S., Lee, H. J., Zhang, Z. and Elisseeff, J. 2010. Regulation of osteogenic and chondrogenic differentiation of mesenchymal stem cells in PEG–ECM hydrogels. Cell Tissue Res., 344(3), 499–509.CrossRefGoogle Scholar
Ingber, D. E., Mow, V. C., Butter, D. et al. 2006. Tissue engineering and developmental biology: going biomimetic. Tissue Eng., 12(12), 3265–83.CrossRefGoogle ScholarPubMed
Jongpaiboonkit, L., King, W. J., Lyons, G. E. et al. 2009. Screening for 3D environments that support human mesenchymal stem cell viability using hydrogel arrays. Tissue Eng. Part A, 15(2), 343–53.CrossRefGoogle ScholarPubMed
Kaihara, S., Matsumura, S. and Fisher, J. P. 2009. Cellular responses to degradable cyclic acetal modified PEG hydrogels. J. Biomed. Mater. Res. Part A, 90(3), 863–73.CrossRefGoogle ScholarPubMed
Kloxin, A. M., Benton, J. A. and Anseth, K. S. 2010. In situ elasticity modulation with dynamic substrates to direct cell phenotype. Biomaterials, 31(1), 1–8.CrossRefGoogle ScholarPubMed
Kloxin, A. M., Kasko, A. M., Salinas, C. N. and Anseth, K. S. 2009. Photodegradable hydrogels for dynamic tuning of physical and chemical properties. Science, 324(5923), 59–63.CrossRefGoogle ScholarPubMed
Kraehenbuehl, T. P., Zammaretti, P., Van der Vlies, A. J. 2008. Three-dimensional extracellular matrix-directed cardioprogenitor differentiation: systematic modulation of a synthetic cell-responsive PEG-hydrogel. Biomaterials, 29(18), 2757–66.CrossRefGoogle ScholarPubMed
Lampe, K. J., Mooney, R. G., Bjugstad, K. B. and Mahoney, M. J. 2010. Effect of macromer weight percent on neural cell growth in 2D and 3D nondegradable PEG hydrogel culture. J. Biomed. Mater. Res. Part A, 94(4), 1162–71.Google ScholarPubMed
Lee, H. J., Lee, J. S., Chansakul, T. et al. 2006. Collagen mimetic peptide-conjugated photopolymerizable PEG hydrogel. Biomaterials, 27(30), 5268–76.CrossRefGoogle ScholarPubMed
Lee, H. J., Yu, C., Chansakul, T. et al. 2008a. Enhanced chondrogenesis of mesenchymal stem cells in collagen mimetic peptide-mediated microenvironment. Tissue Eng. Part A, 14(11), 1843–51.CrossRefGoogle ScholarPubMed
Lee, J., Cuddihy, M. J. and Kotov, N. A. 2008b. Three-dimensional cell culture matrices: state of the art. Tissue Eng. Part B – Rev., 14(1), 61–86.CrossRefGoogle ScholarPubMed
Lee, J. S. and Murphy, W. L. 2010. Modular peptides promote human mesenchymal stem cell differentiation on biomaterial surfaces. Acta Biomater., 6(1), 21–8.CrossRefGoogle ScholarPubMed
Lee, J. S., Lee, J. S., Wagoner-Johnson, A. and Murphy, W. L. 2009. Modular peptide growth factors for substrate-mediated stem cell differentiation. Angewandte Chem. Int. Edition Engl., 48(34), 6266–9.CrossRefGoogle ScholarPubMed
Lee, S. T., Yun, J. I., Jo, Y. S. et al. 2010. Engineering integrin signaling for promoting embryonic stem cell self-renewal in a precisely defined niche. Biomaterials, 31(6), 1219–26.CrossRefGoogle Scholar
Leslie-Barbick, J. E., Moon, J. J. and West, J. L. 2009. Covalently-immobilized vascular endothelial growth factor promotes endothelial cell tubulogenesis in poly(ethylene glycol) diacrylate hydrogels. J. Biomater. Sci. – Polymer Edition, 20(12), 1763–79.CrossRefGoogle ScholarPubMed
Leslie-Barbick, J. E., Saik, J. E., Gould, D. J., Dickinson, M. E. and West, J. L. 2011a. The promotion of microvasculature formation in poly(ethylene glycol) diacrylate hydrogels by an immobilized VEGF-mimetic peptide. Biomaterials, 32(25), 5782–9.CrossRefGoogle ScholarPubMed
Leslie-Barbick, J. E., Shen, C., Chen, C. S. and West, J. 2011b. Micron-scale spatially patterned, covalently immobilized vascular endothelial growth factor on hydrogels accelerates endothelial tubulogenesis and increases cellular angiogenic responses. Tissue Eng. Part A, 17(1–2), 221–9.CrossRefGoogle ScholarPubMed
Lin, C. C. and Anseth, K. S. 2009. Controlling affinity binding with peptide-functionalized poly(ethylene glycol) hydrogels. Adv. Functional Mater., 19(14), 2325–31.CrossRefGoogle ScholarPubMed
Lin, C. C., Boyer, P. D., Aimetti, A. A. and Anseth, K. S. 2010. Regulating MCP-1 diffusion in affinity hydrogels for enhancing immuno-isolation. J. Controlled Release, 142(3), 384–91.CrossRefGoogle ScholarPubMed
Lin, C. C., Metters, A. T. and Anseth, K. S. 2009. Functional PEG–peptide hydrogels to modulate local inflammation induced by the pro-inflammatory cytokine TNF alpha. Biomaterials, 30(28), 4907–14.CrossRefGoogle Scholar
Liu, H. and Roy, K. 2005. Biomimetic three-dimensional cultures significantly increase hematopoietic differentiation efficacy of embryonic stem cells. Tissue Eng., 11(1–2), 319–30.CrossRefGoogle ScholarPubMed
Liu, S. Q., Tian, Q. A., Hedrick, J. L. et al. 2010. Biomimetic hydrogels for chondrogenic differentiation of human mesenchymal stem cells to neocartilage. Biomaterials, 31(28), 7298–307.CrossRefGoogle ScholarPubMed
Lund, A. W., Yener, B., Stegemann, J. P. and Plopper, G. E. 2009. The natural and engineered 3D microenvironment as a regulatory cue during stem cell fate determination. Tissue Eng. Part B – Rev., 15(3), 371–80.CrossRefGoogle ScholarPubMed
Luo, Y. and Shoichet, M. S. 2004. A photolabile hydrogel for guided three-dimensional cell growth and migration. Nature Mater., 3(4), 249–53.CrossRefGoogle ScholarPubMed
Lutolf, M. P. and Hubbell, J. A., 2003. Synthesis and physicochemical characterization of end-linked poly(ethylene glycol)-co-peptide hydrogels formed by Michael-type addition. Biomacromolecules, 4(3), 713–22.CrossRefGoogle ScholarPubMed
Lutolf, M. P. and Hubbell, J. A. 2005. Synthetic biomaterials as instructive extracellular microenvironments for morphogenesis in tissue engineering. Nature Biotechnol., 23(1), 47–55.CrossRefGoogle ScholarPubMed
Lutolf, M. P., Lauer-Fields, J. L., Schmoekel, H. G. et al. 2003a. Synthetic matrix metalloproteinase-sensitive hydrogels for the conduction of tissue regeneration: engineering cell-invasion characteristics. Proc. Nat. Acad. Sci. USA, 100(9), 5413–18.CrossRefGoogle ScholarPubMed
Lutolf, M. P., Raeber, G. P., Zisch, A. H., Tirelli, N. and Hubbell, J. A. 2003b. Cell-responsive synthetic hydrogels. Adv. Mater. 15(11), 888–92.CrossRefGoogle Scholar
Lynn, A. D., Blakney, A. K., Kyriakides, T. R. and Bryant, S. J. 2011. Temporal progression of the host response to implanted poly(ethylene glycol)-based hydrogels. J. Biomed. Mater. Res. Part A, 96(4), 621–31.CrossRefGoogle ScholarPubMed
Lynn, A. D., Kyriakides, T. R. and Bryant, S. J. 2010. Characterization of the in vitro macrophage response and in vivo host response to poly(ethylene glycol)-based hydrogels. J. Biomed. Mater. Res. Part A, 93(3), 941–53.Google ScholarPubMed
Mason, M. N., Arnold, C. A. and Mahoney, M. J. 2009. Entrapped collagen type 1 promotes differentiation of embryonic pancreatic precursor cells into glucose-responsive β-cells when cultured in three-dimensional PEG hydrogels. Tissue Eng. Part A, 15(12), 3799–808.CrossRefGoogle ScholarPubMed
Mason, M. N. and Mahoney, M. J. 2010. Inhibition of γ-secretase activity promotes differentiation of embryonic pancreatic precursor cells into functional islet-like clusters in poly(ethylene glycol) hydrogel culture. Tissue Eng. Part A, 16(8), 2593–603.CrossRefGoogle ScholarPubMed
Maynard, H. D. and Hubbell, J. A. 2005. Discovery of a sulfated tetrapeptide that binds to vascular endothelial growth factor. Acta Biomater., 1(4), 451–9.CrossRefGoogle ScholarPubMed
Moon, J. J., Lee, S. H. and West, J. L. 2007. Synthetic biomimetic hydrogels incorporated with Ephrin-A1 for therapeutic angiogenesis. Biomacromolecules, 8(1), 42–9.CrossRefGoogle ScholarPubMed
Moon, J. J., Saik, J. E., Poché, R. A. et al. 2010. Biomimetic hydrogels with pro-angiogenic properties. Biomaterials, 31(14), 3840–7.CrossRefGoogle ScholarPubMed
Nguyen, L. H., Kudva, A. K., Saxena, N. S. and Roy, K. 2011. Engineering articular cartilage with spatially-varying matrix composition and mechanical properties from a single stem cell population using a multi-layered hydrogel. Biomaterials, 32(29), 6946–52.CrossRefGoogle ScholarPubMed
Nie, T., Baldwin, A., Yamaguchi, N. and Kiick, K. L. 2007. Production of heparin-functionalized hydrogels for the development of responsive and controlled growth factor delivery systems. J. Controlled Release 122(3): 287–296.CrossRefGoogle ScholarPubMed
Nuttelman, C. R., Rice, M. A., Rydholm, A. E. et al. 2008. Macromolecular monomers for the synthesis of hydrogel niches and their application in cell encapsulation and tissue engineering. Prog. Polymer Sci., 33(2), 167–79.CrossRefGoogle ScholarPubMed
Nuttelman, C. R., Tripodi, M. C. and Anseth, K. S. 2005. Synthetic hydrogel niches that promote hMSC viability. Matrix Biol., 24(3), 208–18.CrossRefGoogle ScholarPubMed
Nuttelman, C. R., Tripodi, M. C. and Anseth, K. S. 2006. Dexamethasone-functionalized gels induce osteogenic differentiation of encapsulated hMSCs. J. Biomed. Mater. Res. Part A, 76(1), 183–95.CrossRefGoogle ScholarPubMed
Pierschbacher, M. D. and Ruoslahti, E. 1984. Cell attachment activity of fibronectin can be duplicated by small synthetic fragments of the molecule. Nature, 309, 30–3.CrossRefGoogle ScholarPubMed
Polizzotti, B. D., Fairbanks, B. D. and Anseth, K. S. 2008. Three-dimensional biochemical patterning of click-based composite hydrogels via thiolene photopolymerization. Biomacromolecules, 9(4), 1084–7.CrossRefGoogle ScholarPubMed
Rimmer, S., Tattersall, P., Ebdon, J. R. and Fullwood, N. J. 1999. New strategies for the synthesis of amphiphilic networks. Reactive Functional Polymers, 41(1–3), 177–84.CrossRefGoogle Scholar
Rizzi, S. C., Ehrbar, M., Halstenberg, S. et al. 2006. Recombinant protein-co-PEG networks as cell-adhesive and proteolytically degradable hydrogel matrixes. Part II: biofunctional characteristics. Biomacromolecules, 7(11), 3019–29.CrossRefGoogle ScholarPubMed
Rydholm, A. E., Bowman, C. N. and Anseth, K. S. 2005. Degradable thiol–acrylate photopolymers: polymerization and degradation behavior of an in situ forming biomaterial. Biomaterials, 26(22), 4495–506.CrossRefGoogle Scholar
Saik, J. E., Gould, D. J., Keswani, A. H., Dickinson, M. E. and West, J. L. 2011a. Biomimetic hydrogels with immobilized ephrinA1 for therapeutic angiogenesis. Biomacromolecules, 12(7), 2715–22.CrossRefGoogle ScholarPubMed
Saik, J. E., Gould, D. J., Watkins, E. M., Dickinson, M. E. and West, J. L. 2011b. Covalently immobilized platelet-derived growth factor-BB promotes antiogenesis in biomimetic poly(ethylene glycol) hydrogels. Acta Biomater., 7(1), 133–43.CrossRefGoogle Scholar
Sakiyama-Elbert, S. E. and Hubbell, J. A. 2000a. Controlled release of nerve growth factor from a heparin-containing fibrin-based cell ingrowth matrix. J. Controlled Release, 69(1), 149–58.CrossRefGoogle ScholarPubMed
Sakiyama-Elbert, S. E. and Hubbell, J. A. 2000b. Development of fibrin derivatives for controlled release of heparin-binding growth factors. J. Controlled Release, 65(3), 389–402.CrossRefGoogle ScholarPubMed
Salinas, C. N. and Anseth, K. S. 2008a. Mixed mode thiol–acrylate photopolymerizations for the synthesis of PEG–peptide hydrogels. Macromolecules, 41(16), 6019–26.CrossRefGoogle Scholar
Salinas, C. N. and Anseth, K. S. 2008b. The enhancement of chondrogenic differentiation of human mesenchymal stem cells by enzymatically regulated RGD functionalities. Biomaterials, 29(15), 2370–7.CrossRefGoogle ScholarPubMed
Salinas, C. N. and Anseth, K. S. 2008c. The influence of the RGD peptide motif and its contextual presentation in PEG gels on human mesenchymal stem cell viability. J. Tissue Eng. Regen. Med., 2(5), 296–304.CrossRefGoogle ScholarPubMed
Shapira-Schweitzer, K., Habib, M., Gepstein, L. and Seliktar, D. 2009. A photopolymerizable hydrogel for 3-D culture of human embryonic stem cell-derived cardiomyocytes and rat neonatal cardiac cells. J. Molec. Cellular Cardiol., 46(2), 213–24.CrossRefGoogle ScholarPubMed
Shapira-Schweitzer, K. and Seliktar, D. 2007. Matrix stiffness affects spontaneous contraction of cardiomyocytes cultured within a PEGylated fibrinogen biomaterial. Acta Biomater., 3(1), 33–41.CrossRefGoogle ScholarPubMed
Su, J., Hu, B. H., Lowe, W. L., Kaufman, D. B. and Messersmith, P. B. 2010. Anti-inflammatory peptide-functionalized hydrogels for insulin-secreting cell encapsulation. Biomaterials, 31(2), 308–14.CrossRefGoogle ScholarPubMed
Terella, A., Mariner, P., Brown, N., Anseth, K. and Streubel, S. O. 2010. Repair of a calvarial defect with biofactor and stem cell-embedded polyethylene glycol scaffold. Arch. Facial Plastic Surg., 12(3), 166–71.CrossRefGoogle ScholarPubMed
Tibbitt, M. W. and Anseth, K. S. 2009. Hydrogels as extracellular matrix mimics for 3D cell culture. Biotechnol. Bioeng., 103(4), 655–63.CrossRefGoogle ScholarPubMed
Weber, L. M., Hayda, K. N., Haskins, K. and Anseth, K. S. 2008. Cell–matrix interactions improve β-cell survival and insulin secretion in three-dimensional culture. Tissue Eng. Part A, 14(12), 1959–68.CrossRefGoogle ScholarPubMed
Wei, H. -L., Yang, Z., Chen, Y. et al. 2010. Characterisation of N-vinyl-2-pyrrolidone-based hydrogels prepared by a Diels–Alder click reaction in water. Eur. Polymer J., 46(5), 1032–9.CrossRefGoogle Scholar
Willerth, S. M., Johnson, P. J., Maxwell, D. J. et al. 2007. Rationally designed peptides for controlled release of nerve growth factor from fibrin matrices. J. Biomed. Mater. Res. Part A, 80(1), 13–23.CrossRefGoogle ScholarPubMed
Yamaguchi, N., Zhang, L., Chae, B.-S. et al. 2007. Growth factor mediated assembly of cell receptor-responsive hydrogels. J. Am. Chem. Soc., 129(11), 3040–1.CrossRefGoogle ScholarPubMed
Zieris, A., Prokoph, S., Levental, K. R. et al. 2010. FGF-2 and VEGF functionalization of starPEG-heparin hydrogels to modulate biomolecular and physical cues of angiogenesis. Biomaterials, 31(31), 7985–94.CrossRefGoogle ScholarPubMed

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

Available formats
×