Skip to main content
Log in

Super-Resolution from Noisy Data

  • Published:
Journal of Fourier Analysis and Applications Aims and scope Submit manuscript

Abstract

This paper studies the recovery of a superposition of point sources from noisy bandlimited data. In the fewest possible words, we only have information about the spectrum of an object in the low-frequency band [−f lo,f lo] and seek to obtain a higher resolution estimate by extrapolating the spectrum up to a frequency f hi>f lo. We show that as long as the sources are separated by 2/f lo, solving a simple convex program produces a stable estimate in the sense that the approximation error between the higher-resolution reconstruction and the truth is proportional to the noise level times the square of the super-resolution factor (SRF) f hi/f lo.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

Notes

  1. To be precise, the theorem holds for any feasible point \(\tilde{x}\) obeying \(\| \tilde {x} \| _{\mathrm{TV}}\leq\| x \| _{\mathrm{TV}}\); this set is not empty since it contains x.

References

  1. Batenkov, D., Yomdin, Y.: On the accuracy of solving confluent Prony systems. SIAM J. Appl. Math. 73(1), 134–154 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  2. Betzig, E., Patterson, G.H., Sougrat, R., Lindwasser, O.W., Olenych, S., Bonifacino, J.S., Davidson, M.W., Lippincott-Schwartz, J., Hess, H.F.: Imaging intracellular fluorescent proteins at nanometer resolution. Science 313(5793), 1642–1645 (2006)

    Article  Google Scholar 

  3. Bhaskar, B.N., Tang, G., Recht, B.: Atomic norm denoising with applications to line spectral estimation. Preprint

  4. Bienvenu, G.: Influence of the spatial coherence of the background noise on high resolution passive methods. In: Proceedings of the International Conference on Acoustics, Speech and Signal Processing, vol. 4, pp. 306–309 (1979)

    Google Scholar 

  5. Blu, T., Dragotti, P., Vetterli, M., Marziliano, P., Coulot, L.: Sparse sampling of signal innovations. IEEE Signal Process. Mag. 25(2), 31–40 (2008)

    Article  Google Scholar 

  6. Candès, E.J., Fernandez-Granda, C.: Towards a mathematical theory of super-resolution. Commun. Pure Appl. Math. (2013, to appear). doi:10.1002/cpa.21455

  7. Chi, Y., Scharf, L.L., Pezeshki, A., Calderbank, A.R.: Sensitivity to basis mismatch in compressed sensing. IEEE Trans. Signal Process. 59(5), 2182–2195 (2011)

    Article  MathSciNet  Google Scholar 

  8. Clergeot, H., Tressens, S., Ouamri, A.: Performance of high resolution frequencies estimation methods compared to the Cramér-Rao bounds. IEEE Trans. Acoust. Speech Signal Process. 37(11), 1703–1720 (1989)

    Article  MATH  Google Scholar 

  9. Donoho, D.L.: Superresolution via sparsity constraints. SIAM J. Math. Anal. 23(5), 1309–1331 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  10. Duarte, M.F., Baraniuk, R.G.: Spectral compressive sensing. Appl. Comput. Harmon. Anal. 35(1), 111–129 (2013)

    Article  MathSciNet  Google Scholar 

  11. Dumitrescu, B.: Positive Trigonometric Polynomials and Signal Processing Applications. Springer, Berlin (2007)

    MATH  Google Scholar 

  12. Fannjiang, A., Liao, W.: Coherence-pattern guided compressive sensing with unresolved grids. SIAM J. Imaging Sci. 5, 179 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  13. Fannjiang, A.C.: The MUSIC algorithm for sparse objects: a compressed sensing analysis. Inverse Probl. 27(3), 035013 (2011)

    Article  MathSciNet  Google Scholar 

  14. Grant, M., Boyd, S.: CVX: Matlab software for disciplined convex programming, version 1.21 (2011). http://cvxr.com/cvx

  15. Greenspan, H.: Super-resolution in medical imaging. Comput. J. 52, 43–63 (2009)

    Article  Google Scholar 

  16. Harris, T.D., Grober, R.D., Trautman, J.K., Betzig, E.: Super-resolution imaging spectroscopy. Appl. Spectrosc. 48(1), 14–21 (1994)

    Article  Google Scholar 

  17. Helstrom, C.: The detection and resolution of optical signals. IEEE Trans. Inf. Theory 10(4), 275–287 (1964)

    Article  Google Scholar 

  18. Hess, S.T., Girirajan, T.P., Mason, M.D.: Ultra-high resolution imaging by fluorescence photoactivation localization microscopy. Biophys. J. 91(11), 4258 (2006)

    Article  Google Scholar 

  19. Hu, L., Shi, Z., Zhou, J., Fu, Q.: Compressed sensing of complex sinusoids: an approach based on dictionary refinement. IEEE Trans. Signal Process. 60(7), 3809–3822 (2012)

    Article  MathSciNet  Google Scholar 

  20. Huang, B., Bates, M., Zhuang, X.: Super-resolution fluorescence microscopy. Annu. Rev. Biochem. 78, 993–1016 (2009)

    Article  Google Scholar 

  21. Itakura, F.: Line spectrum representation of linear predictor coefficients of speech signals. J. Acoust. Soc. Am. 57(S1), S35 (1975)

    Article  Google Scholar 

  22. Kennedy, J., Israel, O., Frenkel, A., Bar-Shalom, R., Azhari, H.: Super-resolution in PET imaging. IEEE Trans. Med. Imaging 25(2), 137–147 (2006)

    Article  Google Scholar 

  23. Khaidukov, V., Landa, E., Moser, T.J.: Diffraction imaging by focusing-defocusing: an outlook on seismic superresolution. Geophysics 69(6), 1478–1490 (2004)

    Article  Google Scholar 

  24. Laurent, B., Massart, P.: Adaptive estimation of a quadratic functional by model selection. Ann. Stat. 28(5), 1053–1302 (1992)

    MathSciNet  Google Scholar 

  25. Lindberg, J.: Mathematical concepts of optical superresolution. J. Opt. 14(8), 083001 (2012)

    Article  Google Scholar 

  26. Makovoz, D., Marleau, F.R.: Point source extraction with MOPEX. Publ. Astron. Soc. Pac. 117(836), 1113–1128 (2005)

    Article  Google Scholar 

  27. McCutchen, C.W.: Superresolution in microscopy and the Abbe resolution limit. J. Opt. Soc. Am. 57(10), 1190 (1967)

    Article  Google Scholar 

  28. Milanfar, P. (ed.) Super-Resolution Imaging. Digital Imaging and Computer Vision (2010)

  29. Odendaal, J., Barnard, E., Pistorius, C.: Two-dimensional superresolution radar imaging using the MUSIC algorithm. IEEE Trans. Antennas Propag. 42(10), 1386–1391 (1994)

    Article  Google Scholar 

  30. Park, S.C., Park, M.K., Kang, M.G.: Super-resolution image reconstruction: a technical overview. IEEE Signal Process. Mag. 20(3), 21–36 (2003)

    Article  Google Scholar 

  31. Prony, R.: Essai expérimental et analytique: sur les lois de la dilatabilité de fluides élastique et sur celles de la force expansive de la vapeur de l’alkool, à différentes températures. J. Éc. Polytech. 1(2), 24–76 (1795)

    Google Scholar 

  32. Puschmann, K.G., Kneer, F.: On super-resolution in astronomical imaging. Astron. Astrophys. 436, 373–378 (2005)

    Article  Google Scholar 

  33. Roy, R., Kailath, T.: ESPRIT—estimation of signal parameters via rotational invariance techniques. IEEE Trans. Acoust. Speech Signal Process. 37(7), 984–995 (1989)

    Article  Google Scholar 

  34. Rudin, W.: Real and Complex Analysis, 3rd edn. McGraw-Hill, New York (1987)

    MATH  Google Scholar 

  35. Rust, M.J., Bates, M., Zhuang, X.: Sub-diffraction-limit imaging by stochastic optical reconstruction microscopy (STORM). Nat. Methods 3(10), 793–796 (2006)

    Article  Google Scholar 

  36. Schmidt, R.: Multiple emitter location and signal parameter estimation. IEEE Trans. Antennas Propag. 34(3), 276–280 (1986)

    Article  Google Scholar 

  37. Shahram, M., Milanfar, P.: Imaging below the diffraction limit: a statistical analysis. IEEE Trans. Image Process. 13(5), 677–689 (2004)

    Article  Google Scholar 

  38. Shahram, M., Milanfar, P.: On the resolvability of sinusoids with nearby frequencies in the presence of noise. IEEE Trans. Signal Process. 53(7), 2579–2588 (2005)

    Article  MathSciNet  Google Scholar 

  39. Slepian, D.: Prolate spheroidal wave functions, Fourier analysis, and uncertainty. V. The discrete case. Bell Syst. Tech. J. 57, 1371–1430 (1978)

    Article  MATH  Google Scholar 

  40. Stoica, P., Babu, P.: Sparse estimation of spectral lines: Grid selection problems and their solutions. IEEE Trans. Signal Process. 60(2), 962–967 (2012)

    Article  MathSciNet  Google Scholar 

  41. Stoica, P., Moses, R., Friedlander, B., Soderstrom, T.: Maximum likelihood estimation of the parameters of multiple sinusoids from noisy measurements. IEEE Trans. Acoust. Speech Signal Process. 37(3), 378–392 (1989)

    Article  Google Scholar 

  42. Stoica, P., Moses, R.L.: Spectral Analysis of Signals. Prentice Hall, New York (2005)

    MATH  Google Scholar 

  43. Stoica, P., Nehorai, A.: Statistical analysis of two nonlinear least-squares estimators of sine-wave parameters in the colored-noise case. Circuits Syst. Signal Process. 8(1), 3–15 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  44. Stoica, P., Soderstrom, T.: Statistical analysis of MUSIC and subspace rotation estimates of sinusoidal frequencies. IEEE Trans. Signal Process. 39(8), 1836–1847 (1991)

    Article  MATH  Google Scholar 

  45. Tang, G., Bhaskar, B.N., Shah, P., Recht, B.:. Compressed sensing off the grid. Preprint

Download references

Acknowledgements

E.J. Candès is partially supported by AFOSR under grant FA9550-09-1-0643, by ONR under grant N00014-09-1-0258 and by a gift from the Broadcom Foundation. C. Fernandez-Granda is supported by a Fundación Caja Madrid Fellowship. We thank Carlos Sing-Long for useful feedback about an earlier version of the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Carlos Fernandez-Granda.

Additional information

Communicated by Thomas Strohmer.

Appendices

Appendix A: Proof of Lemma 2.5

We use the construction described in Sect. 2 of [6]. In more detail,

$$ q(t) = \sum_{t_k \in T} \alpha_k G(t-t_k) + \beta_k G^{ ( 1 ) }(t-t_k), $$

where \(\alpha, \beta\in\mathbb{C}^{\vert T \vert }\) are coefficient vectors,

$$ G(t) = \biggl[\frac{\sin ( ( \frac{f_{\mathrm{lo}}}{2}+1 ) \pi t ) }{ ( \frac{f_{\mathrm{lo}}}{2}+1 ) \sin ( \pi t ) } \biggr]^4, \quad t \in\mathbb{T} \setminus\{0\}, $$
(A.1)

and G(0)=1; here, G () is the th derivative of G. If f lo is even, G(t) is the square of the Fejér kernel. By construction, the coefficients α and β are selected such that for all t j T,

$$\begin{aligned} q(t_j) & = v_j \\ q'(t_j) & = 0. \end{aligned}$$

Without loss of generality we consider t j =0 and bound q(t)−v j in the interval [0,0.16λ lo]. To ease notation, we define w(t)=q(t)−v j =w R (t)+iw I (t), where w R is the real part of w and w I the imaginary part. Leveraging different results from Sect. 2 in [6] (in particular the equations in (2.25) and Lemmas 2.2 and 2.7), we have

$$\begin{aligned} \bigl \vert w_R'' ( t ) \bigr \vert & = \biggl \vert \sum_{t_k \in T}\operatorname{Re} ( \alpha_k ) G ^{ ( 2 ) } ( t-t_k ) + \sum _{t_k \in T}\operatorname{Re} ( \beta_k ) G ^{ ( 3 ) } ( t-t_k ) \biggr \vert \\ & \leq\| \alpha \| _{L_\infty}\sum_{t_k \in T}\bigl \vert G^{ ( 2 ) } ( t-t_k ) \bigr \vert + \| \beta \| _{L_\infty}\sum_{t_k \in T} \bigl \vert G^{ ( 3 ) } ( t-t_k ) \bigr \vert \\ & \leq C_{\alpha} \biggl( \bigl \vert G^{ ( 2 ) } ( t ) \bigr \vert + \sum_{t_k \in T\setminus\{0\}} \bigl \vert G^{ ( 2 ) } ( t-t_k ) \bigr \vert \biggr) \\ &\quad\ \ {}+ C_{\beta}\lambda_{\mathrm{lo}} \biggl( \bigl \vert G^{ ( 3 ) } ( t ) \bigr \vert + \sum _{t_k \in T\setminus\{0\}} \bigl \vert G^{ ( 3 ) } ( t-t_k ) \bigr \vert \biggr) \\ & \leq C f_{\mathrm{lo}}^2. \end{aligned}$$

The same bound holds for w I . Since w R (0), \(w_{R}'(0)\), w I (0) and \(w_{I}'(0)\) are all equal to zero, this implies \(\vert w_{R}(t) \vert \leq C' f_{\mathrm{lo}}^{2} t^{2}\) and \(\vert w_{I}(t) \vert \leq C' f_{\mathrm {lo}}^{2} t^{2}\) in the interval of interest, which allows the conclusion

$$\begin{aligned} \bigl \vert w(t) \bigr \vert \leq C f_{\mathrm{lo}}^2 t^2. \end{aligned}$$

Appendix B: Proof of Lemma 2.7

The proof is similar to that of Lemma 2.4 (see Sect. 2 of [6]), where a low-frequency kernel and its derivative are used to interpolate an arbitrary sign pattern on a support satisfying the minimum-distance condition. More precisely, we set

$$ q_1(t) = \sum_{t_k \in T} \alpha_k G(t-t_k) + \beta_k G ^{ ( 1 ) }(t-t_k), $$
(B.1)

where \(\alpha, \beta\in\mathbb{C}^{\vert T \vert }\) are coefficient vectors, G is defined by (A.1). Note that G, G (1) and, consequently, q 1 are trigonometric polynomials of degree at most f 0. By Lemma 2.7 in [6], it holds that for any t 0T and \(t \in\mathbb{T}\) obeying |tt 0|≤0.16λ lo,

$$\begin{aligned} \sum_{t_k \in T\setminus\{t_0\}} \bigl \vert G^{ ( \ell ) } ( t-t_k ) \bigr \vert \leq C_{\ell} f_{\mathrm{lo}}^{\ell}, \end{aligned}$$
(B.2)

where C is a positive constant for =0,1,2,3; in particular, C 0≤0.007, C 1≤0.08 and C 2≤1.06. In addition, there exist other positive constants \(C_{0}'\) and \(C_{1}'\), such that for all t 0T and \(t\in\mathbb{T}\) with |tt 0|≤Δ/2,

$$\begin{aligned} \sum_{t_k \in T\setminus\{t_0\}} \bigl \vert G^{ ( \ell ) } ( t-t_k ) \bigr \vert \leq C_{\ell}' f_{\mathrm{lo}}^{\ell} \end{aligned}$$
(B.3)

for =0,1. We refer to Sect. 2.3 in [6] for a detailed description of how to compute these bounds.

In order to satisfy (2.18) and (2.19), we constrain q 1 as follows: for each t j T,

$$\begin{aligned} q_1(t_j) & = 0, \\ q_1'(t_j) & = v_j. \end{aligned}$$

Intuitively, this forces q 1 to approximate the linear function v j (tt j ) around t j . These constraints can be expressed in matrix form,

$$\left [\begin{array}{c@{\quad}c} D_0 & D_1\\ D_1 & D_2 \end{array} \right ] \begin{bmatrix} \alpha\\ \beta \end{bmatrix} = \begin{bmatrix} 0\\ v \end{bmatrix} , $$

where

$$( D_0 ) _{jk} = G ( t_j - t_k ) , \quad\quad ( D_1 ) _{jk} = G^{ ( 1 ) } ( t_j - t_k ) , \quad\quad ( D_2 ) _{jk} = G ^{ ( 2 ) } ( t_j - t_k ) , $$

and j and k range from 1 to |T|. It is shown in Sect. 2.3.1 of [6] that under the minimum-separation condition this system is invertible, so that α and β are well defined. These coefficient vectors can consequently be expressed as

$$\begin{bmatrix} \alpha\\ \beta \end{bmatrix} = \begin{bmatrix} -D_0^{-1}D_1 \\ {I} \end{bmatrix} S^{-1} v,\quad \quad S := D_2-D_1D_0^{-1}D_1, $$

where S is the Schur complement. Inequality (B.2) implies

$$\begin{aligned} \| {I}- D_0 \| _{\infty} & \le C_0 , \end{aligned}$$
(B.4)
$$\begin{aligned} \| D_1 \| _{\infty} & \le C_{1} f_{\mathrm {lo}}, \end{aligned}$$
(B.5)
$$\begin{aligned} \| \kappa{I}-D_2 \| _{\infty} & \le C_{2} f_{\mathrm{lo}}^2 , \end{aligned}$$
(B.6)

where κ=|G (2)(0)|=π 2 f lo(f lo+4)/3.

Let ∥M denote the usual infinity norm of a matrix M defined as \(\|M\|_{\infty}= \max_{\|x\|_{\infty}= 1} \|Mx\|_{\infty}= \max_{i} \sum_{j} |a_{ij}|\). Then, if ∥IM<1, the series M −1=(I−(IM))−1=∑ k≥0(IM)k is convergent and we have

$$ \| M^{-1} \| _{\infty} \leq\frac {1}{1-\| {I}-M \| _{\infty}}. $$

This, together with (B.4), (B.5) and (B.6) implies

$$\begin{aligned} \| D_0^{-1} \| _{\infty} & \leq\frac {1}{1-\| {I}-D_0 \| _{\infty}} \leq \frac{1}{1-C_0} , \\ \| \kappa{I}- S \| _{\infty} & \leq \| \kappa{I} -D_2 \| _{\infty}+\| D_1 \| _{\infty}\| D_0^{-1} \| _{\infty}\| D_1 \| _{\infty} \leq \biggl( C_2+\frac{C_1^2}{1-C_0} \biggr) f_{\mathrm{lo}}^2 , \\ \| S^{-1} \| _{\infty} & = \kappa^{-1} \biggl\| \biggl( \frac {S}{\kappa} \biggr) ^{-1} \biggr\|_{\infty} \leq\frac{1}{\kappa-\| \kappa{I}-S \| _{\infty}} \leq \biggl( \kappa- \biggl( C_2+\frac{C_1^2}{1-C_0} \biggr) f_{\mathrm{lo}}^2 \biggr) ^{-1} \\ &\leq C_{\kappa} \lambda_{\mathrm{lo}}^{2}, \end{aligned}$$

for a certain positive constant C κ . Note that due to the numeric upper bounds on the constants in (B.2) C κ is indeed a positive constant as long as f lo≥1. Finally, we obtain a bound on the magnitude of the entries of α

$$\begin{aligned} \| \alpha \| _{\infty} & = \| D_0^{-1}D_1 S^{-1} v \| _{\infty} \leq \| D_0^{-1}D_1 S^{-1} \| _{\infty} \leq \| D_0^{-1} \| _{\infty} \| D_1 \| _{\infty} \| S^{-1} \| _{\infty} \leq C_{\alpha} \lambda_{\mathrm{lo}} , \end{aligned}$$
(B.7)

where C α =C κ C 1/(1−C 0), and on the entries of β

$$ \| \beta \| _{\infty}=\| S^{-1} v \| _{\infty}\leq \| S^{-1} \| _{\infty} \leq C_{\beta}\lambda_{\mathrm{lo}}^{2}, $$
(B.8)

for a positive constant C β =C κ . Combining these inequalities with (B.3) and the fact that the absolute values of G(t) and G (1)(t) are bounded by one and 7f lo respectively (see the proof of Lemma C.5 in [6]), we have that for any t

$$\begin{aligned} |q_1(t)| & = \biggl \vert \sum_{t_k \in T} \alpha_k G ( t-t_k ) + \sum_{t_k \in T} \beta_k G^{ ( 1 ) } ( t-t_k ) \biggr \vert \\ & \leq\| \alpha \| _{\infty}\sum_{t_k \in T}\bigl \vert G ( t-t_k ) \bigr \vert + \| \beta \| _{\infty}\sum _{t_k \in T} \bigl \vert G^{ ( 1 ) } ( t-t_k ) \bigr \vert \\ & \leq C_{\alpha} \lambda_{\mathrm{lo}} \biggl( \bigl \vert G ( t ) \bigr \vert + \sum_{t_k \in T\setminus\{t_i\}} \bigl \vert G ( t-t_k ) \bigr \vert \biggr) \\ &\quad\ \ {}+ C_{\beta }\lambda_{\mathrm{lo}}^2 \biggl( \bigl \vert G^{ ( 1 ) } ( t ) \bigr \vert + \sum _{t_k \in T\setminus\{t_i\}} \bigl \vert G^{ ( 1 ) } ( t-t_k ) \bigr \vert \biggr) \\ & \leq C \lambda_{\mathrm{lo}}, \end{aligned}$$
(B.9)

where t i denotes the element in T nearest to t (note that all other elements are at least Δ/2 away). Thus, (2.19) holds.

The proof is completed by the following lemma, which proves (2.18).

Lemma 6.1

For any t j T and \(t \in\mathbb{T}\) obeying |tt j |≤0.16λ lo, we have

$$\begin{aligned} |q_1(t)-v_j ( t-t_j ) | & \leq \frac{C ( t-t_j ) ^2}{\lambda_{{\mathrm{lo}}}}. \end{aligned}$$

Proof

We assume without loss of generality that t j =0. By symmetry, it suffices to show the claim for t∈(0,0.16λ lo]. To ease notation, we define w(t)=v j tq 1(t)=w R (t)+iw I (t), where w R is the real part of w and w I the imaginary part. Leveraging (B.7), (B.8) and (B.2) together with the fact that G (2)(t) and G (3)(t) are bounded by \(4 f_{\mathrm{lo}}^{2}\) and \(6 f_{\mathrm{lo}}^{3}\) respectively if |t|≤0.16λ lo (see the proof of Lemma 2.3 in [6]), we obtain

$$\begin{aligned} \bigl \vert w_R'' ( t ) \bigr \vert & = \biggl \vert \sum_{t_k \in T}\operatorname{Re} ( \alpha_k ) G ^{ ( 2 ) } ( t-t_k ) + \sum _{t_k \in T}\operatorname{Re} ( \beta_k ) G ^{ ( 3 ) } ( t-t_k ) \biggr \vert \\ & \leq\| \alpha \| _{\infty}\sum_{t_k \in T}\bigl \vert G^{ ( 2 ) } ( t-t_k ) \bigr \vert + \| \beta \| _{\infty}\sum_{t_k \in T} \bigl \vert G ^{ ( 3 ) } ( t-t_k ) \bigr \vert \\ & \leq C_{\alpha} \lambda_{\mathrm{lo}} \biggl( \bigl \vert G^{ ( 2 ) } ( t ) \bigr \vert + \sum_{t_k \in T\setminus\{0\} } \bigl \vert G^{ ( 2 ) } ( t-t_k ) \bigr \vert \biggr) \\ &\quad\ \ {} + C_{\beta}\lambda_{\mathrm{lo}}^2 \biggl( \bigl \vert G^{ ( 3 ) } ( t ) \bigr \vert + \sum_{t_k \in T\setminus\{0\}} \bigl \vert G^{ ( 3 ) } ( t-t_k ) \bigr \vert \biggr) \\ & \leq C f_{\mathrm{lo}}. \end{aligned}$$

The same bound applies to w I . Since w R (0), \(w_{R}'(0)\), w I (0) and \(w_{I}'(0)\) are all equal to zero, this implies |w R (t)|≤Cf lo t 2—and similarly for |w I (t)|—in the interval of interest. Whence, |w(t)|≤Cf lo t 2. □

Appendix C: Proof of Corollary 1.3

The proof of Theorem 1.2 relies on two identities

$$\begin{aligned} \| x_{\mathrm{est}} \| _{\mathrm{TV}} & \leq \| x \| _{\mathrm{TV}}, \end{aligned}$$
(C.1)
$$\begin{aligned} \| Q_{\mathrm{lo}} ( x_{\mathrm{est}}- x ) \| _{L_1} & \leq2\delta, \end{aligned}$$
(C.2)

which suffice to establish

$$ \| K_{\mathrm{hi}}*(x_{\mathrm{est}}- x) \| _{L_1} \leq C_0 \mathrm{SRF}^2 \delta. $$

To prove the corollary, we show that (C.1) and (C.2) hold. Due to the fact that \(\| \epsilon \| _{2}^{2}\) follows a χ 2-distribution with 4f lo+2 degrees of freedom, we have

$$ \mathbb{P} \bigl( \| \epsilon \| _{2} > ( 1+\gamma ) \sigma\sqrt{4 f_{\mathrm{lo}}+2}=\delta \bigr) < e^{- 2f_{\mathrm{lo}}\gamma^2 }, $$

for any positive γ by a concentration inequality (see [24, Sect. 4]). By Parseval, this implies that with high probability \(\| Q_{\mathrm{lo}}x -y \| _{\mathcal {L}_{2}}=\| \epsilon \| _{2}\leq\delta\). As a result, x est is feasible, which implies (C.1) and furthermore

$$\begin{aligned} \| Q_{\mathrm{lo}} ( x_{\mathrm{est}}- x ) \| _{L_1} \leq\| Q_{\mathrm{lo}} ( x_{\mathrm{est}} - x ) \| _{\mathcal{L}_2} \leq \| Q_{\mathrm{lo}}x -y \| _{\mathcal {L}_2}+\| y-Q_{\mathrm{lo}}x_{\mathrm{est}} \| _{\mathcal{L}_2} \leq2\delta, \end{aligned}$$

since by the Cauchy-Schwarz inequality \(\| f \| _{L_{1}} \leq\| f \| _{\mathcal{L}_{2}}\) for any function f with bounded L 2 norm supported on the unit interval. Thus, (C.2) also holds and the proof is complete.

Appendix D: Extension to Multiple Dimensions

The extension of the proof hinges on establishing versions of Lemmas 2.4, 2.5 and 2.7 for multiple dimensions. These lemmas construct bounded low-frequency polynomials which interpolate a sign pattern on a well-separated set of points S and have bounded second derivatives in a neighborhood of S. In the multidimensional case, we need the directional derivative of the polynomials to be bounded in any direction, which can be ensured by bounding the eigenvalues of their Hessian matrix evaluated on the support of the signal. To construct such polynomials one can proceed in a way similar to the proof of Lemmas 2.4 and 2.7, namely, by using a low-frequency kernel constructed by tensorizing several squared Fejér kernels to interpolate the sign pattern, while constraining the first-order derivatives to either vanish or have a fixed value. As in the one-dimensional case, one can set up a system of equations and prove that it is well conditioned using the rapid decay of the interpolation kernel away from the origin. Finally, one can verify that the construction satisfies the required conditions by exploiting the fact that the interpolation kernel and its derivatives are locally quadratic and rapidly decaying. This is spelled out in the proof of Proposition C.1 in [6] to prove a version of Lemma 2.4 in two dimensions. In order to clarify further how to adapt our techniques to a multidimensional setting we provide below a sketch of the proof of the analog of Lemma 2.1 in two dimensions. In particular, this illustrates how the increase in dimension does not change the exponent of the SRF in our recovery guarantees.

4.1 D.1 Proof of an Extension of Lemma 2.1 to Two Dimensions

We now have \(t \in\mathbb{T}^{2}\). As a result, we redefine

$$\begin{aligned} S_{\mathrm{near}}^{\lambda} ( j ) & := \{ t : \| t-t_j \| _{L_\infty}\leq w \lambda \}, \\ S_{\mathrm{far}}^{\lambda} &:= \{ t : \| t-t_j \| _{L_\infty}> w \lambda, \forall t_j \in T \}, \\ I_{S_{\mathrm{near}}^{\lambda} ( j ) } ( \mu ) &:= \frac{1}{\lambda_{\mathrm{lo}}^2}\int_{S_{\mathrm{near}}^{\lambda } ( j ) } \| t-t_j \| _{2}^2 \vert \mu \vert ( d t ) , \end{aligned}$$

where w is a constant.

The proof relies on the existence of a low-frequency polynomial

$$ q(t) = \sum_{k_1 = -f_{{\mathrm{lo}}}}^{f_{{\mathrm {lo}}}}\sum _{k_2 = -f_{{\mathrm{lo}}}}^{f_{{\mathrm{lo}}}} c_{k_1,k_2} e^{i2\pi(k_1 t_1 +k_2 t_2)} $$

satisfying

$$\begin{aligned} q(t_j) & = v_j, \quad t_j \in T, \end{aligned}$$
(D.1)
$$\begin{aligned} |q(t)| & \leq1-\frac{C'_a \| t-t_j \| _{2}^2}{\lambda_{{\mathrm{lo}}}^2} , \quad t \in S_{\mathrm{near}}^{\lambda_{{\mathrm{lo}}}} ( j ) , \end{aligned}$$
(D.2)
$$\begin{aligned} |q(t)| & < 1-C'_b , \quad t \in S_{\mathrm{far}}^{\lambda_{{\mathrm{lo}}}}, \end{aligned}$$
(D.3)

where \(C_{a}'\) and \(C_{b}'\) are constants. Proposition C.1 in [6] constructs such a polynomial. Under a minimum distance condition, which constrains the elements of T to be separated by 2.38λ lo in infinity norm (as explained in [6] this choice of norm is arbitrary and could be changed to the 2 norm), [6] shows that q satisfies (D.1) and (D.3) and that both eigenvalues of its Hessian matrix evaluated on T are of order \(f_{\mathrm{lo}}^{2}\), which implies (D.2).

As in one dimension, we perform a polar decomposition of P T h,

$$\begin{aligned} P_{T} h = e^{i \phi ( t ) }\vert P_{T} h \vert , \end{aligned}$$

and work with \(v_{j} = e^{-i \phi(t_{j})}\). The rest of the proof is almost identical to the 1D case. Since q is low frequency,

$$\begin{aligned} \biggl \vert \int_{\mathbb{T}^2} q(t) dh (t) \biggr \vert & \leq2 \delta. \end{aligned}$$
(D.4)

Next, since q interpolates e (t) on T,

$$\begin{aligned} \| P_{T} h \| _{\mathrm{TV}} = \int_{\mathbb {T}^2} q(t) P_{T} h ( dt ) &\leq2\delta+ \sum_{j\in T} \biggl \vert \int_{S_{\mathrm{near}}^{\lambda_{\mathrm{lo}}} ( j ) \setminus \{ t_j \} } q(t) h ( dt ) \biggr \vert + \biggl \vert \int_{S_{\mathrm{far}}^{\lambda_{\mathrm {lo}}}} q(t) h ( dt ) \biggr \vert . \end{aligned}$$
(D.5)

Applying (D.3) and Hölder’s inequality, we obtain

$$\begin{aligned} \biggl \vert \int_{S_{\mathrm{far}}^{\lambda_{\mathrm{lo}}}} q(t) h ( dt ) \biggr \vert & \leq \bigl( 1-C_b' \bigr) \| P_{S_{\mathrm{far}}^{\lambda_{\mathrm{lo}}}} ( h ) \| _{\mathrm{TV}}. \end{aligned}$$
(D.6)

Setting t j =(0,0) without loss of generality, the triangle inequality and (D.2) yield

$$\begin{aligned} \biggl \vert \int_{S_{\mathrm{near}}^{\lambda_{\mathrm{lo}}} ( j ) \setminus \{ (0,0) \}} q(t) h ( dt ) \biggr \vert & \leq \int_{S_{\mathrm{near}}^{\lambda_{\mathrm{lo}}} ( j ) \setminus \{ (0,0) \}} \vert h \vert ( dt ) - C_a' I_{S_{\mathrm{near}}^{\lambda_{\mathrm {lo}}} ( j ) } ( h ) . \end{aligned}$$
(D.7)

Combining (D.5), (D.6) and (D.7) gives

$$\begin{aligned} \| P_{T} h \| _{\mathrm{TV}} \leq& 2\delta+ \| P_{T^{c}}h \| _{\mathrm{TV}} - C_b' \| P_{S_{\mathrm{far}}^{\lambda_{\mathrm{lo}}}} ( h ) \| _{\mathrm{TV}}- C_a' I_{S_{\mathrm {near}}^{\lambda_{\mathrm{lo}}}} ( h ) \end{aligned}$$

and similarly

$$\| P_{T} h \| _{\mathrm{TV}} \leq2\delta+ \| P_{T^{c}}h \| _{\mathrm{TV}} - w^2 C_a' \mathrm{SRF}^{-2} \| P_{S_{\mathrm{far}}^{\lambda _{\mathrm{hi}}}} ( h ) \| _{\mathrm{TV}}- C_a' I_{S_{\mathrm{near}} ^{\lambda_{\mathrm{hi}}}} ( h ) . $$

By the same argument as in the 1D case, the fact that \(\hat{x}\) has minimal total-variation norm is now sufficient to establish

$$\begin{aligned} C_b' \| P_{S_{\mathrm{far}}^{\lambda_{\mathrm {lo}}}} ( h ) \| _{\mathrm{TV}} + C_a' I_{S_{\mathrm{near}}^{\lambda_{\mathrm{lo}} }} ( h ) \leq2 \delta, \end{aligned}$$

and

$$\begin{aligned} w^2 C_a' \mathrm{SRF}^{-2} \| P_{S_{\mathrm {far}}^{\lambda_{\mathrm{hi}}}} ( h ) \| _{\mathrm{TV}} + C_a' I_{S_{\mathrm{near}}^{\lambda_{\mathrm{hi}}}} ( h ) \leq2 \delta. \end{aligned}$$

Rights and permissions

Reprints and permissions

About this article

Cite this article

Candès, E.J., Fernandez-Granda, C. Super-Resolution from Noisy Data. J Fourier Anal Appl 19, 1229–1254 (2013). https://doi.org/10.1007/s00041-013-9292-3

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00041-013-9292-3

Keywords

Mathematics Subject Classification

Navigation