Вы находитесь на странице: 1из 14

Published June 29, 2009 JCB: FEATURE

Accuracy and precision in quantitative fluorescence


microscopy
Jennifer C. Waters
Harvard Medical School, Department of Cell Biology, Boston, MA 02115

The light microscope has long been used During acquisition of the digital image, accuracy is obvious. Precision is equally
to document the localization of fluores­ the photons that are detected at each important in quantitative fluorescence
cent molecules in cell biology research. pixel are converted to an intensity value microscopy because we are often forced
With advances in digital cameras and that is correlated to, but not equal to, the to make only one measurement (for ex-
the discovery and development of geneti­ number of detected photons (Pawley, ample, one time-point in a live-cell
cally encoded fluorophores, there has 2006c). In fluorescence microscopy, the time-lapse experiment). In addition, we
THE JOURNAL OF CELL BIOLOGY

been a huge increase in the use of fluor­ intensity value of a pixel is related to the are usually measuring biological speci-
escence microscopy to quantify spatial number of fluorophores present at the cor- mens that have some level of natural
and temporal measurements of fluores­ responding area in the specimen. We can variability, so variance seen in measure-
cent molecules in biological specimens.

Downloaded from jcb.rupress.org on June 30, 2009


therefore use digital images to extract ments made on different cells will be
Whether simply comparing the relative
two types of information from fluores- caused by both biological variability
intensities of two fluorescent specimens,
cence microscopy images: (1) spatial, and that which is introduced when mak-
or using advanced techniques like Förster
resonance energy transfer (FRET) or fluor­ which can be used to calculate such ing the measurement. To use a fluores-
escence recovery after photobleaching properties as distances, areas, and veloc- cence microscope and digital detector to
(FRAP), quantitation of fluorescence re­ ities; and (2) intensity, which can be used quantitate spatial and intensity informa-
quires a thorough understanding of the to determine the local concentration of tion from biological specimens, we must
limitations of and proper use of the dif­ fluorophores in a specimen. understand and reduce the sources of in-
ferent components of the imaging system. accuracy and imprecision in these types
Here, I focus on the parameters of digital Accuracy and precision of measurements.
image acquisition that affect the accuracy Every quantitative measurement con-
and precision of quantitative fluores­ tains some amount of error. Error in Signal, background, and
cence microscopy measurements. quantitative fluorescence microscopy noise
measurements may be introduced by the In quantitative fluorescence microscopy,
specimen, the microscope, or the detec- we want to measure the signal coming
What information is tor (Wolf et al., 2007; Joglekar et al., from the fluorophores used to label the
present in a fluorescence 2008). Error shows itself as inaccuracy object of interest in our specimen. For
microscopy digital image? and/or imprecision in measurements. example, consider live cells expressing
Quantitative microscopy measurements Inaccuracy results in the wrong answer. GFP-tubulin in which we wish to mea-
are most often made on digital images. A For example, with an inaccurate pH sure the amount of tubulin polymer. The
digital image is created when the optical meter one might carefully measure the signal we are interested in is the photons
image of the specimen formed by the pH of a basic solution many times, each emitted from GFP bound to tubulin in-
microscope is recorded by a detector time finding the pH to be 2.0. Impreci- corporated into microtubules. We use
(usually a charge-coupled device [CCD] sion, on the other hand, results in vari- the pixel intensity values in the digital
camera [Moomaw, 2007; Spring, 2007] ance in repeated measurements and image to localize the tubulin polymers
or photomultiplier tube [PMT; Art, therefore uncertainty in individual mea- and make conclusions about the quan-
2006]) using a two-dimensional grid of surements. With an imprecise pH meter, tity of microtubules. However, the in-
equally sized pixels. The pixels spatially repeated measurements of a solution tensity values in the digital images of
sample the optical image, such that each with pH 7.0 might have a distribution
pixel represents a defined finite sized ranging from 5.0–9.0, with an average © 2009 Waters  This article is distributed under the
terms of an Attribution–Noncommercial–Share Alike–No
area in a specific location in the specimen. value of 7.0. Although the average value Mirror Sites license for the first six months after the pub-
lication date (see http://www.jcb.org/misc/terms.shtml).
of these repeated measurements is After six months it is available under a Creative Commons
accurate, any individual measurement License (Attribution–Noncommercial–Share Alike 3.0 Un-
ported license, as described at http://creativecommons
Correspondence to jennifer_waters@hms.harvard.edu may be inaccurate. The importance of .org/licenses/by-nc-sa/3.0/).

The Rockefeller University Press  $30.00


J. Cell Biol. Vol. 185 No. 7  1135–1148
www.jcb.org/cgi/doi/10.1083/jcb.200903097 JCB 1135
Published June 29, 2009

Figure 1.  Background fluorescence decreases


precision of fluorescence intensity measure-
ments. (A–C) Wide-field images of 6-µm fluores-
cent beads, all displayed with the same scaling
so relative intensity is evident. All images were
collected with the same microscope (model
TE2000U; Nikon) and the same camera (ORCA-
AG; Hamamatsu Photonics) using a Plan-
Apochromat 60x 1.4 NA oil objective lens (Nikon)
and MetaMorph software. (A) Fluorescent beads
(mounted in PBS) with minimal background fluor­
escence. A 400-ms exposure time was used,
and the maximum intensity value of the beads
is 3,800. Bar = 5 µm. (B) A solution of fluoro­
phore (with the same spectral characteristics as
the fluorophore in the bead, diluted in PBS) was
added to the specimen to increase the back-
ground fluorescence. The exposure time had to
be decreased to 100 ms to get the same maxi-
mum intensity value of the beads, 3,800. (C)
Image B, after background subtraction. Because
a shorter exposure time was used in B, fewer
photons from the beads were collected than in A.
Collecting fewer photons from the object of interest means a higher contribution of Poisson noise, and less precise quantitation of fluorescence intensity
values. Therefore, one should work to remove background fluorescence from the image (see Table I) before background subtraction.

Downloaded from jcb.rupress.org on June 30, 2009


the microtubules represent not only the (Fig. 2, B and C). If the signal is at or photons that can be attributed to Poisson
signal of interest coming from the micro- below the noise level, the variation in noise is determined by the Poisson distri-
tubules, but also background and noise intensity caused by noise will make the bution and is equivalent to the square
(Swedlow et al., 2002; Murray et al., signal indistinguishable from the noise root of the total number of detected pho-
2007; Wolf et al., 2007). in quantitative measurements (Fig. 2 B). tons. This formula applies to the number
Background adds to the signal of As the signal increases relative to the of detected photons, not the arbitrary in-
interest, such that the intensity values in noise level, measurements of the sig- tensity values reported by detectors. De-
the digital image are equal to the signal nal become increasingly more precise tected photons, p, can be calculated from
plus the background (Fig. 1). Background (Fig. 2 C). The precision of quantitative intensity values using the equation
in a digital image of a fluorescent speci- microscopy measurements is therefore
men comes from a variety of sources. In limited by the signal-to-noise ratio (SNR)  
p =  f (imax − o) ×  i − o  ,
our example, monomers of tubulin that of the digital image. SNR affects spatial 



remain in the cytoplasm contribute to measurements as well as intensity mea-


the background, as does the cell culture surements; precise determination of the where f is the full well capacity of the
medium the specimen is mounted in location of a fluorescently labeled object detector, imax is the maximum intensity
which contains phenol red, vitamins, depends on SNR (Fig. 2, D and E) value the detector can produce, i is the
and other components that fluoresce. To (Churchman et al., 2005; Yildiz and intensity value being converted to pho-
accurately and precisely measure the Selvin, 2005; Huang et al., 2008; Manley tons, and o is the detector offset. The de-
signal of interest, background should be et al., 2008). tector values can be obtained from the
reduced as much as possible, and must Poisson noise. One type of technical specification sheets available
be subtracted from measurements (Fig. 1; noise found in fluorescence microscopy on the detector manufacturer’s website.
more on this later). digital images comes from the signal we Poisson noise cannot be reduced or
Noise causes variance in the inten- are trying to measure. Measurements of eliminated. However, as the number of
sity values above and below the “real” stochastic quantum events, such as num- counted signal photons increases, the
intensity value of the signal plus the bers of photons, are fundamentally lim- Poisson noise becomes a smaller per-
background. The extent of deviation dif- ited by Poisson statistics (Pawley, 1994, centage of the signal and the SNR in-
fers from one pixel to the next in a single 2006a). This means that the number of creases. Working to increase the number
digital image, with the maximum vari- photons counted in repeated measure- of signal photons collected will therefore
ance in an image referred to as the noise ments of an ideal, unchanging specimen increase the accuracy and precision of
level (Fig. 2 A). Noise causes impreci- will have a Poisson distribution. The quantitative measurements.
sion in measurements of pixel intensity number of photons counted in a single Maximizing signal. The inten-
values, and therefore a level of uncer- measurement therefore has an intrinsic sity of the signal in digital fluorescence
tainty in the accuracy of the measure- statistical uncertainty called Poisson microscopy images is affected by every
ments. To detect the presence of a signal, noise (also referred to as shot noise, sig- step along the path to quantitation, includ-
the signal must be significantly higher nal noise, or photon noise). The maxi- ing specimen preparation, the microscope,
than the noise level of the digital image mum variance in the number of counted and the detector (Table I, Table II).

 1136 JCB • VOLUME 185 • NUMBER 7 • 2009


Published June 29, 2009 JCB: FEATURE

Figure 2.  The importance of SNR in intensity and spatial measurements. (A) A digital image taken with a cooled CCD camera (ORCA-AG; Hamamatsu
Photonics), with no light sent to the camera. Using MetaMorph software, a line (shown in red) was drawn across the bead and a line-scan graph was gener-
ated to show the intensity value of the pixels along the line. The graph shows line-scans of two similar images, taken in quick succession. The intensity values
in the images fluctuate (range, 195–205) around the camera digital offset value of 200. Notice that the fluctuation in intensity values changes at each pixel
from one image to the next. This variance is due primarily to thermal and readout noise from the CCD camera, and the extent of the variance will differ

Downloaded from jcb.rupress.org on June 30, 2009


depending on the camera. This type of noise is superimposed on every fluorescence microscopy image. (B–E) Images of 6-µm beads that are fluorescently
stained along their perimeter were collected with a wide-field microscope (model TE2000U; Nikon) using a Plan-Apochromat 60x 1.4 NA oil objective
lens, the same camera as in A (ORCA-AG; Hamamatsu Photonics), and MetaMorph software. Line-scans generated as described for A. (B) An image of
the bead taken with a 100-ms exposure time. The SNR is very low, making the bead indistinguishable from the noise in the line scan. (C) An image of the
same bead as in B, taken with a longer (3 s) exposure time. The high SNR of this image would make quantitation of the intensity of the bead, or localization
of the edge of the bead, highly precise. (D and E) The same bead images in B and C. Two images of the bead were taken, and one copy pseudo-colored
red and one copy pseudo-colored green. The pseudo-colored images were shifted relative to one another by a few pixels and merged. (D) With low SNR
images, it is nearly impossible to precisely locate the edges of the beads. (E) With high SNR images, the intensity line scan can be fit to Gaussian curves
and the center located with nanometer precision. This allows the distance between objects of two wavelengths to be precisely determined, even if it is well
below the resolution limit of the microscope (Churchman et al., 2005; Yildiz and Selvin, 2005; Huang et al., 2008; Manley et al., 2008). Bar = 5 µm.

The specimen. Fluorophores Giepmans et al., 2006; Johnson, 2006) excite the fluorophore and collecting as
vary greatly in their intrinsic brightness to make the best choice of fluorophore much of the emission light as possible
and the rate at which they photobleach; and anti-photobleaching reagent for your (Ploem, 1999; Rietdorf and Stelzer, 2006).
an easy way to maximize signal is to specimen and experiment. Goodwin (2007) There are several useful online tools avail-
choose a brighter and more photo-stable provides a complete discussion of the able for matching fluorophores to filters
fluorophore (Diaspro et al., 2006; Tsien importance of mounting medium choice (for example, as of the date of this publi-
et al., 2006). The brightness of a fluoro- to both signal intensity and resolution. cation Invitrogen has a very useful tool
phore is determined primarily by its ex- The microscope. To get the on their website: http://www.invitrogen
tinction coefficient and quantum yield, brightest signal while minimizing speci- .com/site/us/en/home/support/Research-
properties that are dependent on the fluor­ men damage, it is important to use illu- Tools/Fluorescence-SpectraViewer.html).
ophore’s environment (Diaspro et al., mination wavelengths that will optimally In an epifluorescence microscope,
2006). It should be noted that new fluor­ excite the fluorophore and to collect as the objective lens both illuminates the
escent proteins are routinely introduced many of the emission photons as possi- specimen and collects photons emitted
that outperform their predecessors; it is ble (Ploem, 1999; Rietdorf and Stelzer, from fluorophores to form the optical
therefore advisable to search the current 2006). Fluorescence spectra that show image. The numerical aperture (NA) of
scientific literature for the latest variants. the absorption and emission efficiency of the objective lens (marked on the barrel
Fixed specimens should be mounted fluorophores are available from the man- of the lens after the magnification; Keller,
in a glycerol-based mounting medium ufacturer or in the scientific literature 2006) is an important determinant of the
(Egner and Hell, 2006; Goodwin, 2007) (for example, see Shaner et al., 2005), brightness of the optical image. The num-
that contains an anti-photobleaching in- and filter manufacturers provide spectra ber of photons an objective can collect
hibitor (Diaspro et al., 2006). No one anti- online that show the percent transmis- from a specimen (and therefore the bright-
photobleaching reagent is the best, as sion of their filters across wavelength. It ness of the image) increases with NA2.
each reagent is more or less effective for is important to compare the spectra for Brightness of an objective is also deter-
a given fluorophore (Diaspro et al., 2006). the fluorophore you are imaging to spec- mined by properties such as transmission
Review the fluorophore manufacturer’s tra for the fluorescence filter sets (and/or and correction for aberration (Keller,
product information or the relevant scien- laser illumination line) to ensure you are 2006). Spherical aberration caused by the
tific literature (Shaner et al., 2005; using the correct wavelengths of light to objective lens (Hell and Stelzer, 1995;

QUANTITATIVE MICROSCOPY • Waters 1137


Published June 29, 2009

Table I. Checklist for optimizing images for quantitation


Increase signal:

 Choose a bright (high quantum yield, high extinction coefficient) and photo-stable fluorophorea
 Image through a clean No. 1.5 coverslipb
 Mount specimen as close to the coverslip as possiblec
 Use high NA clean objective lens with lowest acceptable magnificationd
 Choose fluorescence filter sets that match fluorophore spectrad
 Align arc lamp for Koehler illuminationd
 For fixed specimens, use a glycerol-based mounting medium containing anti-photobleaching inhibitors3
 Remove DIC Wollaston prism and analyzer from light pathe
 Use a cooled CCD camera with at least 60% quantum efficiencyd
 Use camera binningd

Decrease noise:
 Use a cooled CCD camera with less than 8 electrons readout noise and negligible dark noisef
 Use amplification (e.g., EM-CCDs) only when signal is limitingf
 Increase signal (see above) to reduce relative contribution of Poisson noisef

Decrease background:
 Clean coverslips and opticse
 Perfect fluorophore labeling protocol to minimize nonspecific labelingg
 Mount specimens in minimally fluorescent medium (e.g., without phenol red)d
 Use band-pass filter sets that block autofluorescenced

Downloaded from jcb.rupress.org on June 30, 2009


 Turn off the room lightsd
 Close down the field diaphragm to illuminate only the object of interestd
 When out-of-focus fluorescence is high, consider using deconvolution, confocal, or TIRFh
a
(Diaspro et al., 2006)
b
(Keller, 2006)
c
(Goodwin, 2007)
d
(Waters, 2007)
e
(Inoué and Spring, 1997)
f
(Moomaw, 2007)
g
(Allan, 2000)
h
(Murray, 1998)

Goodwin, 2007) or introduced by the the immersion medium (e.g., mounting Increasing the exposure time al-
specimen (Egner and Hell, 2006) decreases medium with a high concentration of lows the flux of photons coming from the
image intensity (North, 2006; Waters, glycerol will have a refractive index specimen to accumulate (as electrons)
2007; Waters and Swedlow, 2007). Spher- close to that of standard immersion oil). in the detector, increasing the intensity
ical aberration occurs when there is a rel- The detector. The number of values in the image—up to a point
atively large difference in refractive index photons reaching the detector that are (Moomaw, 2007; Spring, 2007; Waters,
between the specimen and the lens im- collected and contribute to the intensity 2007). Detectors have a limited capacity
mersion medium; for example, when an values in a digital image depends on the to hold electrons; if this capacity is
oil immersion lens is used to image a quantum efficiency (QE) of the detec- reached, the corresponding pixel will be
specimen in an aqueous solution such as tor, and how long the signal is allowed “saturated” and any photons reaching the
cell culture medium (Egner and Hell, to integrate on the detector (usually re- detector after saturation will not be
2006). Spherical aberration caused by ferred to as the exposure time). QE is a counted. The linearity of the detector is
refractive index mismatch generally in- measure of the percentage of photons therefore lost, and saturated images can-
creases with distance from the coverslip reaching the detector that are counted not be used for quantitation of fluores-
(Joglekar et al., 2008). Spherical aber- (Moomaw, 2007). The QE of research- cence intensity values. Choosing to “crop
ration can be addressed using water im- grade CCD cameras most often used for out” saturated areas is not acceptable
mersion objective lenses (Keller, 2006), by quantitation of fluorescence images (unless they can be shown to be irrele-
using an objective lens with a correction ranges from 60% to over 90%, whereas vant to the experimental hypothesis) be-
collar (Keller, 2006; Waters, 2007), or by the QE of PMTs used in point-scanning cause it will select for the weaker
immersion oil refractive index matching confocals is much lower, usually 10– intensity parts of the specimen. Satura-
(Goodwin, 2007). For fixed specimens, 20% (although the effective QE is sig- tion should be avoided by using image
spherical aberration is reduced by mount- nificantly less; see Pawley, 2006b). QE acquisition software to monitor intensity
ing fixed specimens in a mounting medium values are available online from the de- values when setting up the acquisition
with a refractive index similar to that of tector manufacturer. parameters (Table II).

 1138 JCB • VOLUME 185 • NUMBER 7 • 2009


Published June 29, 2009 JCB: FEATURE

Table II. Protocol for quantitation of fluorescence intensity values


1. Acquire optical images

• Set up specimen and imaging system for optimal signal detection, low background, and low noise (Table I)

2. Acquire digital images


• Use software to monitor intensity values in the image to choose the best acquisition settingsa
• Use full dynamic range of the camera for fixed specimensa
• For live-cell work, it is often necessary to sacrifice SNR to minimize specimen exposure to light and maintain cell health and viabilitya
• Consider binning to increase SNRa
• Avoid high camera gain when a large dynamic range is neededa
• Avoid saturating pixels in the imagea
• Eliminate or minimize exposure of specimen to fluorescence excitation light prior to image acquisitiona
• Focus carefully, preferably with phase or DICb

3. Store images
• Always save the raw imagesc
• Use either no compression or lossless compressionc

4. Process images
• Use flat-field correction to correct for uneven illuminationd
• Be sure any other image processing used prior to quantitation preserves relative intensity valuesc,d

5. Analyze images
• Subtract local background value from intensity measurementse

Downloaded from jcb.rupress.org on June 30, 2009


• Do not measure intensity values on compressed or pseudo-colored imagesc
• Validate image segmentation and analysis methodf
• Calculate and report the error in your measurementsd,g
a
(Waters, 2007)
b
(Inoué and Spring, 1997)
c
(Russ, 2007)
d
(Wolf et al., 2007)
e
(Hoffman et al., 2001)
f
(Dorn et al., 2008)
g
(Cumming et al., 2007)

In most live biological specimens, ments of intensity, it is also very important Poisson noise. Subtracting a constant
saturation is much less of a problem than to first reduce background as much as background value from intensity measure-
collecting enough signal to get adequate possible (Fig. 1, Table I). Background in ments does not change the variance due to
SNR images for quantitation. Many an image effectively reduces both the dy- Poisson noise; the presence of background
research-grade cooled cameras allow bin- namic range and the SNR. Dynamic range therefore reduces image SNR.
ning of adjacent pixels on the CCD chip. of a CCD camera is defined as the full A common source of background
With all other acquisition parameters be- well capacity of the photodiodes (i.e., the in biological specimens is out-of-focus
ing equal, binning on the CCD chip in- number of photons that can be detected fluorescence. In fluorescence micros-
creases the intensity of the pixels without per pixel before saturation) divided by the copy, the illuminating light is focused at
increasing readout noise, resulting in a detector noise (Moomaw, 2007; Spring, the image focal plane by the objective
higher SNR digital image (Moomaw, 2007). High dynamic range is particularly lens, such that maximum excitation of
2007; Spring, 2007; Waters, 2007). How- important for collecting an adequate num- fluorophores occurs at the focal plane
ever, because the resulting pixels repre- ber of signal photons from both dim and (Hiraoka et al., 1990). However, illumi-
sent a larger area of the specimen (i.e., 4x bright areas of the specimen. Photons nating light above and below the image
larger with a 2 × 2 bin), binning decreases from background sources fill the detector, focal plane excites fluorophores above
the resolution of the digital image (Fig. 3). limiting the number of signal photons that and below the image focal plane. Light
In many low-light imaging experiments, can be collected before the detector satu- emitted from these out-of-focus fluoro-
however, the decrease in resolution is well rates (Fig. 1) and effectively decreasing phores is collected by the objective lens,
worth the increase in SNR (Table II). dynamic range. In addition, recall that the and appears as out-of-focus background
Background fluorescence re- number of photons counted defines the in the in-focus image of the specimen. In
duces dynamic range and de- Poisson noise level in an image. Poisson wide-field epifluorescence microscopy,
creases SNR. Although it’s true that noise is equal to the square root of the sig- adjust­ing the diameter of the field dia-
background fluorescence can and must be nal photons plus background photons; phragm to match the visible field of view
subtracted from quantitative measure- higher background therefore means higher minimizes the illumination of out-of-focus

QUANTITATIVE MICROSCOPY • Waters 1139


Published June 29, 2009

Figure 3.  Resolution and sampling. (A–C) Im-


ages of the same pair of 150-nm green fluores-
cent beads collected with a microscope (model
TE2000U; Nikon), a Plan-Apochromat 100x
1.4 NA oil objective lens, and MetaMorph
software. A camera with 6.45-µm photodiodes
(ORCA-AG; Hamamatsu Photonics) was used,
and different camera binning settings were used
to vary the area of the specimen covered by one
pixel. Exposure times were adjusted to reach a
maximum intensity value of 3,600 for each
image. Using the equation for lateral resolution,
we can calculate that the diameter of the first
minimum of the airy disk, and therefore the diameter of the bead in the optical image, should be equal to 465 nm. Bar = 0.5 µm. (A) An image collected
with no camera binning, and an exposure time of 200 ms. Each pixel corresponds to 65 nm of the specimen, and each bead is sampled with 7 pixels.
(B) An image collected using 2 × 2 camera binning, and an exposure time of 50 ms. Each pixel corresponds to 129 nm of the specimen, and each
bead is sampled with about 3.5 pixels. (C) An image collected using 4 × 4 camera binning, and an exposure time of 25 ms. Each pixel corresponds to
258 nm of the specimen, and each bead is sampled with less than 2 pixels. The optical image is under-sampled, and the two beads can no longer be
distinguished as separate from one another.

fluorophores (Hiraoka et al., 1990) and re- methods used to remove out-of-focus fluor­ or j (pixels in the background), obj is the
duces background (Waters, 2007). escence has limitations, and may contrib- object of interest, bkg is the background,
There are several microscopy tech- ute additional noise to the image (Murray and N is the number of pixels in the object
niques that serve to reduce the amount of et al., 2007). In specimens with low levels of interest or the background. This equa-

Downloaded from jcb.rupress.org on June 30, 2009


out-of-focus fluorescence in the image. of out-of-focus fluorescence (which is tion corrects for different-sized regions of
Confocal microscopes illuminate the spec- often the case in adherent cultured cells), interest used to measure the object of in-
imen with a focused light source, while standard wide-field fluorescence micros- terest and the background by calculating
one or more corresponding pinholes at the copy may result in the highest SNR image the background per pixel. This can also
image plane block out-of-focus fluores- (Murray et al., 2007). Therefore, none of be achieved by using image analysis soft-
cence from reaching the detector (Pawley, the different modes of microscopy is ware to calculate the mean intensity value
2006b). Spot-scanning confocals scan the “better” than the other, only more or less in a region of interest, as long as the num-
specimen point-by-point with a single fo- appropriate for a particular specimen or ber of pixels in the region of interest and
cused laser beam, whereas multi-point or application (Swedlow et al., 2002; Murray the range of intensity values in those pix-
slit-scanning confocals (including spin- et al., 2007). When possible, empirical els are sufficient to give a precise mean
ning disk confocals) use multiple pinholes comparison of available modes is the most (Cumming et al., 2007). To avoid errors
or slits to illuminate the specimen more reliable way to ensure you are using the due to an inhomogeneous background, it
quickly (Adams et al., 2003; Tommre and best imaging system for your application. is best to make background measurements
Pawley, 2006). Multi-photon microscopes Any background that remains in a using pixels that are immediately adja-
illuminate the specimen with a focused fluorescence microscopy digital image cent to or surrounding the object of inter-
high-power long wavelength laser, which must be subtracted from intensity value est (for examples, see Hoffman et al.,
results in excitation of the fluorophores measurements to reveal the signal (Table 2001 or Murray et al., 2007). This is espe-
through absorption of multiple photons at II). Consider two specimens, one with an cially important when making measure-
the same time only at the focal plane average intensity value of 2,000 and the ments of intracellular structures because
(Rocheleau and Piston, 2003). In total second with an average intensity value of the background in the cytoplasm is often
internal reflection (TIRF) microscopy, 2,500. Without considering background, different than the background outside of
fluorophores are excited with the evanes- one might conclude that the fluorescence cells, and is usually inhomogeneous.
cent wave of energy that forms when total signal in these two specimens differs by Detector noise. Fluorescence
internal reflection occurs at the boundary 25%. However, if the background in each microscopy digital images are degraded
between media of different refractive in- image measures 1,900, the difference is by Poisson noise and by noise from the
dexes, usually the coverslip and the speci- actually sixfold! Background should be detector (Pawley, 1994, 2006a; Moomaw,
men (Axelrod et al., 1983). Deconvolution subtracted following the equation 2007; Spring, 2007). Thermal noise is
algorithms can also be used to reduce the caused by the stochastic generation of
out-of-focus fluorescence in digital images j = Nbkg thermal electrons within the detector, and

post-acquisition (Wallace et al., 2001).
Because out-of-focus fluorescence i = Nobj ∑ Fbkg j

is largely eliminated by cooling (hence
the use of cooled CCD cameras; Table I).
j =1 ,
is a source of background, and background
Fobj =
∑ Fobji − Nobj
Nbkg Read noise is generated by the amplifier
i =1
reduces SNR and dynamic range, shouldn’t circuitry used to measure the voltage at
we always use one of the imaging methods each pixel, and is usually the dominant
that reduces out-of-focus fluorescence? where F is the fluorescence intensity mea- source of noise in standard cooled CCD
The answer is not that simple. Each of the sured at each pixel i (pixels in the object) cameras designed for quantitative imaging.

 1140 JCB • VOLUME 185 • NUMBER 7 • 2009


Published June 29, 2009 JCB: FEATURE

Read noise is usually expressed in the choice of detector and acquisition set- equations define the theoretical resolu-
manufacturer’s technical specifications tings (Table I). tion limits in most cases; it should be
as a number of electrons, meaning that noted that a handful of very talented
the measured voltage will have a vari- Resolution microscopists have found it possible to
ance equal to that number of electrons In digital microscopy, spatial resolution surpass these limits using specialized
(i.e., the lower the value, the lower the is defined by both the microscope and “super-resolution” imaging techniques (for
noise). Detectors that use signal amplifi- the detector, and limits our ability to ac- review see Evanko, 2009).
cation (e.g., PMTs and electron multi- curately and precisely locate an object It is a common misconception
plying [EM] CCDs) introduce addi­tional and distinguish close objects as separate among cell biologists that confocal mi-
noise during the amplification process. from one another (Inoue, 2006; Rasnik croscopy should be used to obtain the
For example, EM-CCD cameras amplify et al., 2007; Waters and Swedlow, 2007). highest resolution images. Although
signal differences sufficiently to reveal Objects that cannot be detected in an confocal microscopy is very effective at
clock-induced charging—stochastic vari- image cannot be resolved, so spatial increasing contrast in specimens with sig-
ations in the transfer of charge from one resolution is dependent on image SNR nificant out-of-focus fluorescence (Pawley,
pixel to another during read operations (Pawley, 2006c). When imaging a dy- 2006b; Murray et al., 2007), the obtain-
(Robbins and Hadwen, 2003; Moomaw, namic specimen over time, accuracy able resolution of confocal microscopy
2007). When possible, collecting more of quantitation may be further limited is essentially the same as conventional
photons from the specimen to increase by temporal resolution (the rate of wide-field fluorescence microscopy
the signal (see Table I) will result in a image acquisition; see Dorn et al., 2005; (Inoue, 2006). In addition, Murray et al.
higher SNR image than amplifying a Jares-Erijman and Jovin, 2006 for (2007) recently demonstrated that the
smaller number of collected photons. an example). photon collection efficiency and SNR of

Downloaded from jcb.rupress.org on June 30, 2009


The various sources of noise add as the Resolution in the optical wide-field fluorescence is generally higher
sum of the squares: image. Lateral resolution of the optical than confocal microscopy for specimens
image is defined as the distance by with limited out-of-focus fluorescence.
which two objects must be separated in Confocal microscopy becomes necessary
2
Ntotal = N Poisson + N2 + N2 + ... . order to distinguish them as separate and favorable for specimens with high lev-
read thermal
from one another, which is equal to the els of out-of-focus fluorescence because
radius of the smallest point source in the out-of-focus fluorescence adds to back-
The resulting total noise in the digital image (defined as the first minimum of ground fluorescence, and therefore de-
image defines a minimum expected vari- the airy disk; Inoue, 2006). Lateral reso- creases the capacity to collect the signal of
ance in the measured intensity values. lution (r) in epifluorescence microscopy interest (Fig. 1; Murray et al., 2007).
Differences in measurements that lie is given by Resolution in the digital
within the expected variance due to noise image. The optical image is sampled by a
cannot be attributed to the specimen. r = (0.61)λ ,
NA detector to create a digital image. The reso-
Pawley (1994) provides a thorough re- lution of a digital image acquired with a
view of the different sources of noise in where  is the wavelength of emission CCD camera depends on the physical size
digital microscopy images. light and NA is the numerical aperture of of the photodiodes that make up the chip
Noise is not a constant, so it cannot the objective lens. Numerical aperture is (Rasnik et al., 2007), whereas in point-
be subtracted from a digital image. How- usually marked on the objective lens bar- scanning confocal resolution is determined
ever, if multiple images of the same field rel, just after the magnification (Keller, by the area of the specimen that is scanned
of view are collected and averaged to- 2006). Resolution in the z-axis (z) is per pixel (Pawley, 2006c). The pixel size
gether (“frame averaging”), the noise worse than lateral resolution in the light should be at least two times smaller than
will average out and the resulting mean microscope, and is given by the resolution limit of the microscope op-
intensity values will be closer to the tics, so that the smallest possible object in
z = 2λη2 ,
“real” intensity values of the signal plus the image (defined as the diameter of the
 NA
 

the background (Cardullo and Hinch-  


airy disk) will be sampled by 4 pixels
cliffe, 2007). Frame averaging is very where  is the refractive index of the (Fig. 3; Pawley, 2006c). There is a trade-
useful when imaging fixed specimens specimen. It is important to understand off between resolution of the digital image
with a higher noise instrument like a that these equations give the theoretical and signal intensity because magnification
point-scanning confocal, but is usually resolution limits of a perfect lens used to decreases image intensity (Waters, 2007)
impractical for quantitative imaging of image an ideal specimen; real lenses and and smaller pixels generally collect fewer
live fluorescent specimens that are dy- specimens often introduce aberrations in photons (Fig. 3). To compensate for loss of
namic and susceptible to phototoxicity the image that reduce resolution (Egner signal due to smaller pixel size, longer
and photobleaching. For quantitative flu- and Hell, 2006; Goodwin, 2007). The camera exposure times or more intense il-
orescence imaging, noise added to the best way to know the resolution limit of lumination may be necessary (Fig. 3 A). If
digital image during acquisition should your imaging system is to measure it em- the pixel size is too large, the optical image
be reduced as much as possible through pirically (Hiraoka et al., 1990). These will be under-sampled and detail will be

QUANTITATIVE MICROSCOPY • Waters 1141


Published June 29, 2009

Figure 4.  Non-uniform illumination results in nonuniform fluorescence. All images were collected using a microscope (model TE2000E; Nikon), a Plan-
Apochromat 20x 0.75 NA objective lens, a camera (ORCA-AG; Hamamatsu Photonics), and MetaMorph software. (A) An image of a field of fluorescent
beads, using wide-field illumination. Individual beads contain a similar concentration of fluorophore (clumps of beads appear brighter, as is seen near the
center of the image). A pseudo-color displaying the range of intensity values (see inset) was applied. Note that beads in the top left have different intensity
values than the beads in the bottom right. (B) An image of a uniform field of fluorophore taken with the same microscope optics and conditions as A, show-
ing uneven illumination across the field of view. This nonuniform illumination explains the nonuniform fluorescence from the beads of similar fluorophore
concentration shown in A. (C) After flat-field correction (Zwier et al., 2004; Wolf et al., 2007), the image intensity values more accurately reflect the real
fluorescence in the specimen. This image was obtained using the image arithmetic function in image processing software (in this case, MetaMorph) to

Downloaded from jcb.rupress.org on June 30, 2009


divide the image in A by the image in B. Bar = 50 µm.

lost in the digital image (Fig. 3 C). In live- of one object can be accurately deter- indicator dyes, are also reliable for
cell imaging, it is often favorable to give up mined; then intensity values can be used quantitation (Johnson, 2006). Quantita-
some resolution (by binning pixels, for to count multiple objects that are too close tion of live versus fixed cells is gener-
example) to increase image SNR and/or to one another to spatially resolve. These ally preferable because the fixation and
decrease photobleaching and photo dam- types of measurements are very challeng- extraction process can remove tagged
age (Waters, 2007). ing to perform with accuracy, and require proteins, quench the fluorescence of
Optical resolution need not a thorough understanding of, and atten- fluorescent proteins, and change the
limit accuracy in localization or tion to, every possible pitfall (Pawley, size and shape of cells (Allan, 2000;
counting. How does the resolution 2000)—but they are possible. For exam- Straight, 2007).
limit affect our ability to quantitate ple, the measured intensity of proteins One should use caution when using
using fluorescence microscopy? Clearly, conjugated to fluorescent proteins has immunofluorescence to measure the
the size of an object that is below the been used to accurately and precisely local concentration of a protein of inter-
resolution limit cannot be accurately count the number of labeled proteins lo- est, particularly with soluble proteins.
measured with the light microscope. calized to the kinetochore (Joglekar et al., The fixation and extraction necessary to
However, objects that are below the res- 2006) and proteins involved in cytokine- get antibodies into cells can change the
olution can be detected and an image of sis (Wu and Pollard, 2005). quantity and localization of detectable
the object formed by the microscope, if epitopes (Melan and Sluder, 1992). Multi­
the imaging system is sensitive enough Additional threats to valent antibodies also bind with higher
and the object is bright enough. Al- accuracy and precision in affinity to multiple epitopes, which can
though the size of the object in the im- quantitative microscopy make areas with high concentration of
age will be inaccurate, the centroid of a Specimen preparation. Fluorescence the epitope label more efficiently than
high SNR image of the object can be from a fluorophore tagged to a molecule areas of low concentration (Mason and
used to locate the object with nanome- of interest is often used to measure the Williams, 1980). In addition, penetration
ter precision, far beyond the resolution quantity of the tagged molecule. Fluor­ of the antibody may not be consistent in
limit (Churchman et al., 2005; Yildiz escent proteins expressed in live cells different areas of the tissue, cells, and
and Selvin, 2005; Huang et al., 2008; are excellent for quantitation; because subcellular compartments (Allan, 2000).
Manley et al., 2008). there is a constant number of fluoro- Therefore, although accurate quantita-
In fluorescence microscopy, the res- phores per labeled protein, the number tion of the emission photons from fluor­
olution limit does not limit our ability to of photons emitted can be an accurate escently labeled antibodies is possible,
accurately count fluorescently labeled ob- measure of the quantity of fluorescently that number of photons may not accu-
jects, even if the objects are below the labeled protein (Shaner et al., 2005; rately reflect the number of epitopes in
resolution limit. If the objects are all of Straight, 2007; Joglekar et al., 2008). the specimen. It is possible to use rigor-
similar size, are all labeled with the same Small molecules that bind with high ous controls to demonstrate that an im-
number of fluorophores, and the intensity affinity to their target, such as calcium munofluorescence protocol is an accurate

 1142 JCB • VOLUME 185 • NUMBER 7 • 2009


Published June 29, 2009 JCB: FEATURE

Figure 5.  Bleed-through can cause inaccuracy in intensity


measurements. (A and B) Images of a cell (outlined in white)
labeled with DAPI (nuclei) and Bodipy-FL phalloidin (actin).
Both images were collected using the same microscope (model
80i; Nikon), a Plan-Apochromat 100x NA 1.4 oil objective
lens, the same camera (ORCA R2; Hamamatsu Photonics), and
MetaMorph software. The same camera acquisition settings
were used for both images, but they were collected using two
different filters designed for imaging DAPI. (A) An image col-
lected with a DAPI filter set containing a long-pass emission
filter, which allows bleed-through of the Bodipy-FL signal in the
cytoplasm. The bleed-through of the actin in the cytoplasm is
just barely visible by eye in the image. The average intensity
value of the cytoplasm in this image is 205. (B) An image of
the same cell as in A, collected with a DAPI filter set containing
a band-pass emission filter, which blocks bleed-through of the
Bodipy-FL signal in the cytoplasm. The average intensity value
of the cytoplasm in this image is 91, over 50% less than the
image in A. Bar = 10 µm.

measure of a particular epitope of inter- microscope (using a liquid light guide, (Aubin, 1979; Tsien et al., 2006).
est (Mortensen and Larsson, 2001). for example) can increase the uniformity For quantitative measurements, bleed-
Measuring fluorescence deep into of illumination across the field of view through and autofluorescence should
specimens can also be problematic for (Nolte et al., 2006). In many cases, com- be avoided when possible, and measured
measurements of signal intensities. In pletely uniform illumination is impossi- and subtracted from measurements when

Downloaded from jcb.rupress.org on June 30, 2009


biological specimens, light scattering ble to achieve, and one must instead unavoidable (Rietdorf and Stelzer, 2006).
and optical aberrations increase with dis- correct for uneven illumination before Avoid bleed-through and autofluores-
tance from the coverslip and decrease making quantitative measurements (Fig. 4; cence by carefully choosing fluoro-
signal intensity (Murray, 2005). These Table II; Zwier et al., 2004; Wolf et al., phores and filter sets. Bleed-through
effects are difficult to characterize and 2007). To perform the correction, an im- between two fluorophores can be de-
correct for in inhomogeneous biological age of a uniform fluorescent sample is tected and measured by using the fluoro-
samples. Although the effects of light collected to reveal the uneven illumina- phore “A” filter set and camera acquisition
scattering are minimized when using tion pattern (Fig. 4 B). The image to be settings to collect an image of a control
multi-photon illumination (Rocheleau corrected is then divided by the image of sample labeled with only fluorophore
and Piston, 2003), the accuracy of quan- the uniform fluorescent sample (using “B”. Autofluorescence can be detected
titation of intensity values is limited by the image arithmetic functions available and measured by collecting images of a
optical aberrations and dramatically in- in most image-processing software) to control specimen that is identical to the
creased photobleaching in the focal plane obtain the flat-field corrected image experimental specimen except for the
(Patterson and Piston, 2000). (Fig. 4 C). Because the pattern of illumi- addition of exogenous fluorophores.
Non-uniform illumination. nation may differ from day to day, it is Image registration. Images of
Fluorescence emission is generally pro- best to collect a new image of a uniform different wavelengths emitted from a sin-
portional to the intensity of the illumi- fluorescent sample at each imaging ses- gle plane of a specimen may not be coin-
nating light (except when fluorophore sion. A protocol for flat-field correction cident in the optical images (Fig. 6 A).
ground state depletion occurs; see Tsien is described by Wolf et al. (2007). Shifts between wavelengths in X, Y,
et al., 2006; Murray et al., 2007). There- Bleed-through and auto­ and Z can be introduced by a variety of
fore, if a uniform fluorescent sample is fluor­escence. When choosing fluores­ sources, including wedges in fluores-
unevenly illuminated, the resulting fluor­ cence filter sets, maximizing excitation cence filters (Rietdorf and Stelzer, 2006)
escence will usually be uneven as well. and emission collection should be bal- and chromatic aberrations in the objec-
Uneven illumination can be extremely anced with minimizing bleed-through tive lens (Keller, 2006). If registration
detrimental to quantitative measure- (also called crosstalk) and autofluores- between images of different wave-
ments because it may cause the intensity cence. Bleed-through of one fluoro- lengths is important to your quantitative
of an object in one area of the field of phore’s emission through the filter set of analysis, you should check for shifts be-
view to measure differently than the in- another fluorophore can occur when a tween wavelengths in your microscope
tensity of an object of equal fluorophore specimen is labeled with multiple fluoro- using submicron beads infused with
concentration in another area of the field phores whose excitation and emission multiple fluorophores (such as Tetra­
of view (Fig. 4 A). To reduce uneven il- spectra overlap (Fig. 5; Ploem, 1999; speck beads, Invitrogen; protocol in
lumination, the wide-field fluorescence Bolte and Cordelieres, 2006; Rietdorf Wolf et al., 2007). Consistent shifts be-
microscope should be carefully aligned and Stelzer, 2006). Many biological tween wavelengths can then be cor-
for Koehler illumination (Salmon and specimens contain native autofluorescence rected using image-processing software
Canman, 2001). Scrambling the image of similar wavelengths to the emission of (Fig. 6 B) or (for axial shifts) by using a
of the light source before it enters the many commonly used fluorophores focus motor to adjust focus between

QUANTITATIVE MICROSCOPY • Waters 1143


Published June 29, 2009

Downloaded from jcb.rupress.org on June 30, 2009


Figure 6.  Shifts in image registration can affect colocalization results. (A and B) Images of 100-nm Tetra-Speck beads (Invitrogen; mounted in glycerol) that
fluoresce multiple colors including red and green, collected with a microscope (model 80i; Nikon) and camera (ORCA R2; Hamamatsu Photonics) using a
Plan-Apochromat 100x NA 1.4 oil objective lens and MetaMorph software. One image of the beads was collected using a filter set for green fluorescence
(FITC) and a second image of the beads was collected using a filter set for red fluorescence (TRITC); all other microscope optics were the same between
the two images. The two images were pseudo-colored and merged using MetaMorph software. The scatter plots (generated in MetaMorph) display the
correlation between the intensity values of the red and green pixels in the images. (A) The merged image, showing a registration shift of several pixels
between the red and green images. 10 sets of images were collected to determine that the shift is repeatable, and therefore most likely caused by the filter
sets (not depicted). The correlation coefficient for these red and green images is only 0.72, even though the red and green images represent the exact same
beads. Bar = 1 µm. (B) The same images as in A, after correction for the shift in registration using MetaMorph image processing software. The correlation
coefficient increased to 0.97 after the correction.

wavelengths (see Murray et al., 2007 for in 3D is almost always necessary for ac- on the rate of photobleaching, exciting
an example). curately determining intensity of objects fluor­escent specimens before collect-
Focus. Focus is critical to accu- larger than the diffraction limit, and ing images that will be used for
rate and precise quantitation of fluores- when tracking the movements of objects quantitation may introduce error in
cence intensity values (Murray, 2005; that occur in 3D (De Mey et al., 2008). measurements of signal intensity. Ini-
Table II). The distribution of intensity Photobleaching. Almost all tial focusing and scanning of the speci-
values along the z-axis of the optical fluorophores photobleach when exposed men is best done using techniques such
image depends on the size of the fluores- to excitation light in the microscope, as phase or differential interference
cently labeled object and the point spread including all of the fluorescent pro- contrast (DIC) microscopy (Inoué and
function of the microscope. For small teins. The rate of photobleaching is Spring, 1997) because the halogen light
objects imaged with high resolution op- specific to the fluorophore, its environ- sources typically used for transmitted
tics, small changes in focus can have ment, and the intensity of the illuminat- light illumination usually will not bleach
large effects on measured intensity val- ing light (Diaspro et al., 2006). For fluorophores. To accurately measure
ues. Joglekar et al. (2008) describe a some specimens and fluorophores, anti- fluorescence intensity in the same field
method of determining the error intro- photobleaching reagents can be added of view over time, one should measure
duced by imprecise focus when measur- to the mounting medium to reduce the and correct for photobleaching that
ing objects that are thinner than the rate of photobleaching (Diaspro et al., occurs while imaging (Rabut and
diffraction limit of the optics. Measuring 2006; Tsien et al., 2006). Depending Ellenberg, 2005).

 1144 JCB • VOLUME 185 • NUMBER 7 • 2009


Published June 29, 2009 JCB: FEATURE

Image processing and pitfalls to accuracy and precision in quan- can occur only when at least three condi-
storage. Some types of image process- titative microscopy measurements de- tions are met: (1) the emission spectrum
ing and storage can change the relative scribed throughout this review (and in of the donor fluorophore overlaps with
intensity values in a digital image, ren- Pawley, 2000; North, 2006; Wolf et al., the absorption spectrum of the acceptor
dering them unusable for quantitative 2007). In this section, I will address some fluorophore, (2) the donor and acceptor
measurements (Russ, 2007). For example, of the additional specific issues surround- fluorophores are within 10 nm or less of
pseudo-coloring, bit-depth conversion, ing these methods, and will refer the one another, and (3) the emission dipole
and some types of image compression interested reader to more thorough treat- of the donor and the absorption dipole of
(e.g., JPEG) can all compromise the in- ments of each subject. the acceptor are orientated in the correct
tensity values in digital images (Table Colocalization. In its simplest position (i.e., not perpendicular) relative
II). Before using any image-processing and least informative form, colocalization to one another (Stryer, 1978; Schaufele et
algorithm for a quantitative study, be analysis is performed by pseudo-coloring al., 2005). In cell biology research, FRET
sure to understand how it affects image and merging two or more fluorescence experiments usually involve tagging two
intensity values. For example, image- images together, and looking for visual molecules of interest with two different
processing software packages refer to cues that the different wavelengths are fluorophores that are capable of FRET.
many different types of algorithms as present in the same pixels; for example, The presence or absence of FRET is then
“deconvolution”, but not all of these al- yellow pixels in a merged image of red used to make conclusions about the prox-
gorithms are appropriate for quantitation and green fluorophores are often used to imity of the two molecules to which the
(see Wallace et al., 2001). In general, conclude that the two fluorophores “co- different fluorophores are attached. There
analysis of pixel intensity values should localize”. These types of qualitative ob- are many excellent reviews and books
be done on raw images stored without servations show, at best, that both available on quantitative FRET micros-

Downloaded from jcb.rupress.org on June 30, 2009


further scaling or processing, or on im- fluorophores reside within the same 3D copy (Gordon et al., 1998; Siegel et al.,
ages that have been corrected using volume whose minimum size is defined 2000; Jares-Erijman and Jovin, 2003,
methods that have been demonstrated to by the resolution limits of the micro- 2006; Sekar and Periasamy, 2003;
preserve the linear relationship between scope. For high resolution wide-field or Periasamy and Day, 2005; Chen et al.,
photons and image intensity values (e.g., confocal microscopy, this volume is at 2006; Vogel et al., 2006).
flat-field corrected 16-bit TIFF images least one order of magnitude larger than A warning regarding quantitative
are a good choice for quantitation). most large protein complexes. Quantita- FRET microscopy is appropriate. FRET
tive statistical analyses of both the experiments are relatively easy to con-
Common types of spatial distribution and the correlation ceive, but infamously difficult to perform
quantitative microscopy between the intensities of different fluor­ properly. Jares-Erijman and Jovin (2003)
analyses escence channels is a much more infor- provide an overview of the many differ-
Quantitative measurements of spatial and mative way to measure colocalization ent available methods for detecting and
intensity information in fluorescence mi- (Day, 2005; Bolte and Cordelieres, 2006). measuring FRET using fluorescence mi-
croscopy digital images can be used to Bolte and Cordelieres (2006) provide a croscopy. In cell biology research, quan-
answer many different questions about thorough and focused discussion of co- titative measurements of FRET are most
biological specimens. Co-localization, localization analysis using fluorescence commonly performed using wide-field
FRET, and FRAP are three of the most microscopy techniques. or confocal imaging to measure the in-
commonly used quantitative microscopy The accuracy of colocalization tensities of steady-state absorption and
methods in cell biology research. The analyses depends on the ability to discrim- emission of the donor and acceptor mol-
best way to perform these types of exper- inate between the different fluorophores, ecules (Stryer, 1978; Schaufele et al.,
iments depends on many different aspects and on correct registration between images 2005). These methods are plagued by
of the experimental design—for example, of the different wavelengths. Bleed- problems that must be addressed, includ-
the molecule(s) being studied, the through and auto-fluorescence is highly ing auto-fluorescence, noise, photo-
fluorophore(s), the type of specimen, the problematic for colocalization studies, bleaching, and variations in fluorophore
type of microscope, the method of image and must be avoided or corrected for expression level. Spectral bleed-through
analysis, and the hypothesis being tested. (discussed in detail above; Fig. 5; of the fluorophores into the FRET chan-
It is therefore impossible to give a step- Rietdorf and Stelzer, 2006). Registration nel artificially increases the observed
by-step protocol for any of these tech- shifts between wavelengths limit the ac- FRET signal, and must always be mea-
niques that will work for every experiment. curacy of high resolution colocalization sured and corrected for (Sekar and
Even those who are very experienced in measurements, and should be measured Periasamy, 2003; Schaufele et al., 2005;
these techniques must empirically test the and corrected before colocalization analy­ Vogel et al., 2006). Vogel et al. (2006)
experimental design and imaging param- sis (Fig. 6). give an excellent review of the potential
eters for each novel experiment to find FRET. Förster resonance energy problems with accuracy in FRET mea-
the optimal conditions for image acquisi- transfer (FRET) is the nonradiative trans- surements and analysis.
tion and analysis. Each of these methods fer of the energy absorbed by a fluoro- FRAP. In fluorescence recovery
requires careful attention to the various phore to a neighboring fluorophore. FRET after photobleaching (FRAP) experiments,

QUANTITATIVE MICROSCOPY • Waters 1145


Published June 29, 2009

Table III. Information to include in the Materials and methods or figure legend
Follow the journal’s instructions for authors, and/or include the following:

• Manufacturer and model of microscope


• Objective lens magnification, NA, and correction for aberration (e.g., 60x 1.4 NA Plan-Apochromat)
• Fluorescence filter set manufacturer and part number and/or the transmission and bandwidth (e.g., 490/30)
• Illumination light source (including wavelength for laser illumination)
• Camera manufacturer and model
• Software program(s) and version
• Manufacturer and model of other acquisition hardware, including confocal, filter wheels, focus motors, motorized stage, shutters, etc.
• Image acquisition settings including exposure times, gain, and binning
• Other acquisition parameters, including focus step size (for z-series), time between images (for time lapse), etc.
• Description of image processing routine used to create figures
• Description of segmentation and image analysis routine, and method of validation

intense focused illumination is used to of the effect of reversibility of photo- for the experiment (Table III; Waters and
photobleach fluorophores in a select re- bleaching on FRAP measurements de- Swedlow, 2007). Even with this informa-
gion of a specimen. If the fluorescently pends on the rate of image acquisition, tion, it may be difficult for the reader to
labeled molecules are mobile, un- and can be controlled by collecting a assess whether your system was opti-
bleached molecules will move into the number of pre-bleach images to reach a mized and operated to obtain the best

Downloaded from jcb.rupress.org on June 30, 2009


bleached region while the bleached mol- steady state of fluorophores in the dark possible SNR and resolution. Easy open
ecules move out. Recovery of the photo- state (Rabut and Ellenberg, 2005). access to raw image files and data used
bleached region over time can then be for published quantitative analyses will
used to detect mobility/immobility, and Presenting quantitative therefore be critical to the continued
to measure diffusion or association/ microscopy measurements growth of the field of quantitative fluores-
dissociation kinetics of the fluorescently Reporting error. No matter how care- cence microscopy (Moore et al., 2008).
labeled component (Snapp et al., 2003; ful you are when collecting images and
Rabut and Ellenberg, 2005; Sprague and making measurements, every quantitative Further reading
McNally, 2005). analysis has a level of uncertainty that This review is an introduction to the is-
As the specimen is illuminated to must be reported (Table II; Cumming et sues surrounding accurate and precise
collect images of the recovery process, al., 2007; Wolf et al., 2007). Error is most quantitation of fluorescence microscopy
fluorophores will continue to photobleach commonly reported by stating or showing digital images. A thorough appreciation
at a rate dependent (in part) on the level of (as error bars) the standard deviation or of both the power and the limitations of
illumination. To get an accurate measure standard error of the mean of the mea- quantitative microscopy is best obtained
of recovery of the photobleached region, sured values. The appropriate way to re- through careful attention to how these is-
one must measure and correct for photo- port error depends on your data and the sues affect your own data. The interested
bleaching that occurs during image acqui- conclusions you would like to make. This reader is encouraged to learn more from
sition. This can be done, for example, by journal recently published a thorough and the many excellent books, reviews, and
measuring the fluorescence of an un- user-friendly review on reporting error in quantitative analyses that are referenced
bleached region in the same or a neigh- quantitative measurements (Cumming et throughout this review.
boring cell (Rabut and Ellenberg, 2005). al., 2007). When performing arithmetic The author wishes to thank the faculty and students
The accuracy of FRAP analyses is on multiple quantitative measurements of the Analytical and Quantitative Light Micros-
compromised by the fact that fluoro- that have independent sources of error copy course at the Marine Biological Laboratory
(www.mbl.edu/education) for many illuminating
phores can reversibly photobleach (e.g., subtracting an average background discussions—especially Jason Swedlow, David
(Diaspro et al., 2006). It is therefore im- intensity value B ± b from a measurement Wolf, and John Murray; Jason Swedlow and
portant to choose a fluorophore for FRAP of signal intensity S ± s), the error in the Wendy Salmon for helpful comments and discus-
sions on this manuscript; Cassandra Rogers for
experiments that is less likely to enter individual measurements should be prop- preparing bead slides used to acquire images for
illumination-induced dark states that can agated to the final value. A general for- the figures; and Dan Bolton for patient support dur-
spontaneously recover to the fluorescent mula for error propagation was derived by ing the preparation of this manuscript.
state. EGFP, for example, is less likely to Wolf et al. (2007).
undergo reversible photobleaching than Writing the materials and
References
YFP variants. Reversible photobleaching methods. When publishing quantita-
Adams, M.C., W.C. Salmon, S.L. Gupton, C.S.
can be detected by measuring the fluores- tive microscopy data, you should provide Cohan, T. Wittmann, N. Prigozhina, and
cence intensity of entire cells after bleach- the reader with the information they need C.M. Waterman-Storer. 2003. A high-speed
multispectral spinning-disk confocal micro-
ing because it will cause the intensity to to judge whether you used the appropri- scope system for fluorescent speckle micros-
increase within a few seconds. The extent ate equipment and acquisition parameters copy of living cells. Methods. 29:29–41.

 1146 JCB • VOLUME 185 • NUMBER 7 • 2009


Published June 29, 2009 JCB: FEATURE

Allan, V.J. 2000. Protein Localization by Fluorescence In Handbook of Biological Confocal Micros­ Murray, J.M. 1998. Evaluating the performance
Microscopy: A Practical Approach. Oxford copy. J.B. Pawley, editor. Springer-Verlag of fluorescence microscopes. J. Microsc.
University, New York. 256 pp. New York, Inc., New York. 347–353. 191:128–134.
Art, J. 2006. Photon detectors for confocal micros- Hiraoka, Y., J.W. Sedat, and D.A. Agard. 1990. Murray, J.M. 2005. Confocal microscopy, deconvo-
copy. In Handbook of Biological Confocal Determination of three-dimensional imag- lution and structured illumination methods.
Microscopy. J.B. Pawley, editor. Springer- ing properties of a light microscope system. In Live Cell Imaging: A Laboratory Manual.
Verlag New York, Inc., New York. 251–262. Partial confocal behavior in epifluorescence R.D. Goldman and D.L. Spector, editors.
Aubin, J.E. 1979. Autofluorescence of viable microscopy. Biophys. J. 57:325–333. Cold Spring Harbor Laboratory, Cold Spring
cultured mammalian cells. J. Histochem. Hoffman, D.B., C.G. Pearson, T.J. Yen, B.J. Howell, Harbor, NY. 239–279.
Cytochem. 27:36–43. and E.D. Salmon. 2001. Microtubule- Murray, J.M., P.L. Appleton, J.R. Swedlow, and
Axelrod, D., N.L. Thompson, and T.P. Burghardt. dependent changes in assembly of micro- J.C. Waters. 2007. Evaluating performance
1983. Total internal inflection fluorescent tubule motor proteins and mitotic spindle in three-dimensional fluorescence micros-
microscopy. J. Microsc. 129:19–28. checkpoint proteins at PtK1 kinetochores. copy. J. Microsc. 228:390–405.
Mol. Biol. Cell. 12:1995–2009. Nolte, A., J. Pawley, and L. Horing. 2006. Non-laser
Bolte, S., and F.P. Cordelieres. 2006. A guided tour
into subcellular colocalization analysis in Huang, B., W. Wang, M. Bates, and X. Zhuang. light sources for three-dimentional micros-
light microscopy. J. Microsc. 224:213–232. 2008. Three-dimensional super-resolution copy. In Handbook of Biological Confocal
imaging by stochastic optical reconstruction Microscopy. J.B. Pawley, editor. Springer-
Cardullo, R.A., and E.H. Hinchcliffe. 2007. Digital microscopy. Science. 319:810–813. Verlag New York, Inc., New York. 126–144.
manipulation of brightfield and fluorescence
images: noise reduction, contrast enhance- Inoue, S. 2006. Foundations of confocal scanned North, A.J. 2006. Seeing is believing? A beginners’
ment, and feature extraction. Methods Cell imaging in light microscopy. In Handbook guide to practical pitfalls in image acquisition.
Biol. 81:285–314. of Biological Confocal Microscopy. J.B. J. Cell Biol. 172:9–18.
Pawley, editor. Springer-Verlag New York,
Chen, H., H.L. Puhl III, S.V. Koushik, S.S. Vogel, Inc., New York. 1–19. Patterson, G.H., and D.W. Piston. 2000. Photo­
and S.R. Ikeda. 2006. Measurement of bleaching in two-photon excitation micros-
FRET efficiency and ratio of donor to ac- Inoué, S., and K.R. Spring. 1997. Video Microscopy: copy. Biophys. J. 78:2159–2162.
ceptor concentration in living cells. Biophys. The Fundamentals. Plenum Press, New
York. 770 pp. Pawley, J. 2000. The 39 steps: a cautionary tale of
J. 91:L39–L41. quantitative 3-D fluorescence microscopy.
Churchman, L.S., Z. Okten, R.S. Rock, J.F. Dawson, Jares-Erijman, E.A., and T.M. Jovin. 2003. FRET Biotechniques. 28:884–886, 888.
and J.A. Spudich. 2005. Single molecule imaging. Nat. Biotechnol. 21:1387–1395.
Pawley, J.B. 1994. Sources of noise in 3D micro-
Jares-Erijman, E.A., and T.M. Jovin. 2006. Imaging

Downloaded from jcb.rupress.org on June 30, 2009


high-resolution colocalization of Cy3 and scopical data sets. In Three-Dimensional
Cy5 attached to macromolecules measures molecular interactions in living cells by Confocal Microscopy: Volume Investigation
intramolecular distances through time. Proc. FRET microscopy. Curr. Opin. Chem. Biol. of Biological Systems. J.K. Stevens, L.R.
Natl. Acad. Sci. USA. 102:1419–1423. 10:409–416. Mills, and J.E. Trogadis, editors. Academic
Cumming, G., F. Fidler, and D.L. Vaux. 2007. Error Joglekar, A.P., D.C. Bouck, J.N. Molk, K.S. Bloom, Press, Inc., San Diego. 48–90.
bars in experimental biology. J. Cell Biol. and E.D. Salmon. 2006. Molecular architec- Pawley, J.B. 2006a. Fundamental limits in confo-
177:7–11. ture of a kinetochore-microtubule attach- cal microscopy. In Handbook of Biological
ment site. Nat. Cell Biol. 8:581–585. Confocal Microscopy. J.B. Pawley, editor.
Day, R.N. 2005. Imaging protein behavior inside the
living cell. Mol. Cell. Endocrinol. 230:1–6. Joglekar, A.P., E.D. Salmon, K.S. Bloom, and F.S. Springer-Verlag New York, Inc., New York.
Kevin. 2008. Counting kinetochore protein 20–42.
De Mey, J.R., P. Kessler, J. Dompierre, F.P.
numbers in budding yeast using genetically Pawley, J.B. 2006b. Handbook of Biological
Cordelières, A. Dieterlen, J.L. Vonesch, J.B.
encoded fluorescent proteins. Methods Cell Confocal Microscopy. Springer-Verlag New
Sibarita, and F.S. Kevin. 2008. Fast 4D mi-
Biol. 85:127–151. York, Inc., New York. 985 pp.
croscopy. Methods Cell Biol. 85:83–112.
Johnson, I. 2006. Practical considerations in the Pawley, J.B. 2006c. Points, pixels, and gray lev-
Diaspro, A., G. Chirico, C. Usai, P. Ramoino,
selection and application of fluorescent els: digitizing image data. In Handbook
and J. Dobrucki. 2006. Photobleaching.
probes. In Handbook of Biological Confocal of Biological Confocal Microscopy. J.B.
In Handbook of Biological Confocal
Microscopy. J.B. Pawley, editor. Springer- Pawley, editor. Springer-Verlag New York,
Microscopy. J.B. Pawley, editor. Springer-
Verlag New York, Inc., New York. 353–364. Inc., New York. 59–79.
Verlag New York, Inc., New York. 690–699.
Keller, H.E. 2006. Objective lenses for confocal Periasamy, A., and R.N. Day. 2005. Molecular
Dorn, J.F., K. Jaqaman, D.R. Rines, G.S. Jelson,
microscopy. In Handbook of Biological Imaging: FRET Microscopy and Spectroscopy.
P.K. Sorger, and G. Danuser. 2005. Yeast
Confocal Microscopy. J.B. Pawley, editor. Oxford University Press. 336 pp.
kinetochore microtubule dynamics analyzed Springer-Verlag New York, Inc., New York.
by high-resolution three-dimensional mi- 145–160. Ploem, J.S. 1999. Fluorescence microscopy. In
croscopy. Biophys. J. 89:2835–2854. Fluorescent and Luminescent Probes for
Manley, S., J.M. Gillette, G.H. Patterson, H. Shroff, Biological Activity. W.T. Mason, editor.
Dorn, J.F., G. Danuser, G. Yang, and F.S. Kevin. H.F. Hess, E. Betzig, and J. Lippincott-
2008. Computational processing and analy- Academic Press, London. 3–13.
Schwartz. 2008. High-density mapping
sis of dynamic fluorescence image data. of single-molecule trajectories with photo­ Rabut, G., and J. Ellenberg. 2005. Photobleaching
Methods Cell Biol. 85:497–538. activated localization microscopy. Nat. techniques to study mobility and molecular dy-
Egner, A., and S.W. Hell. 2006. Aberrations in Methods. 5:155–157. namics of proteins in live cells: FRAP, iFRAP,
confocal and multi-photon fluorescence mi- and FLIP. In Live Cell Imaging: A Laboratory
Mason, D.W., and A.F. Williams. 1980. The kinetics
croscopy induced by refractive index mis- Manual. R.D. Goldman and D.L. Spector,
of antibody binding to membrane antigens
match. In Handbook of Biological Confocal editors. Cold Spring Harbor Laboratory, Cold
in solution and at the cell surface. Biochem.
Microscopy. J.B. Pawley, editor. Springer- Spring Harbor, NY. 101–126.
J. 187:1–20.
Verlag New York, Inc., New York. 404–412. Rasnik, I., T. French, K. Jacobson, and K. Berland.
Melan, M.A., and G. Sluder. 1992. Redistribution
Evanko, D. 2009. Primer: fluorescence imaging under 2007. Electronic cameras for low-light mi-
and differential extraction of soluble proteins
the diffraction limit. Nat. Methods. 6:19–20. croscopy. Methods Cell Biol. 81:219–249.
in permeabilized cultured cells. Implications
Giepmans, B.N., S.R. Adams, M.H. Ellisman, and for immunofluorescence microscopy. J. Cell Rietdorf, J., and E.H.K. Stelzer. 2006. Special optical
R.Y. Tsien. 2006. The fluorescent toolbox Sci. 101:731–743. elements. In Handbook of Biological Confocal
for assessing protein location and function. Moomaw, B. 2007. Camera technologies for low Microscopy. J.B. Pawley, editor. Springer-
Science. 312:217–224. light imaging: overview and relative advan- Verlag New York, Inc., New York. 43–58.
Goodwin, P.C. 2007. Evaluating optical aberration tages. Methods Cell Biol. 81:251–283. Robbins, M.S., and B.J. Hadwen. 2003. The noise
using fluorescent microspheres: methods, Moore, J., C. Allan, J. Burel, B. Loranger, D. performance of electron multiplying charge-
analysis, and corrective actions. Methods MacDonald, J. Monk, J.R. Swedlow, and coupled devices. IEEE Trans. Electron. Dev.
Cell Biol. 81:397–413. F.S. Kevin. 2008. Open tools for storage 50:1227–1232.
Gordon, G.W., G. Berry, X.H. Liang, B. Levine, and and management of quantitative image data. Rocheleau, J.V., and D.W. Piston. 2003. Two-photon
B. Herman. 1998. Quantitative fluorescence Methods Cell Biol. 85:555–570. excitation microscopy for the study of living
resonance energy transfer measurements Mortensen, K., and L.I. Larsson. 2001. Quantitative cells and tissues. Curr. Protoc. Cell Biol.
using fluorescence microscopy. Biophys. J. and qualitative immunofluorescence studies Chapter 4:Unit 4.11.
74:2702–2713. of neoplastic cells transfected with a con- Russ, J.C. 2007. The Image Processing Handbook.
Hell, S.W., and E. Stelzer. 1995. Lens aberra- struct encoding p53-EGFP. J. Histochem. CRC/Taylor and Francis, Boca Raton, FL.
tions in confocal fluorescence microscopy. Cytochem. 49:1363–1367. 647 pp.

QUANTITATIVE MICROSCOPY • Waters 1147


Published June 29, 2009

Salmon, E.D., and J.C. Canman. 2001. Proper application to molecular motors. Acc. Chem.
alignment and adjustment of the light mi- Res. 38:574–582.
croscope. Curr. Protoc. Cell Biol. Chapter Zwier, J.M., G.J. Van Rooij, J.W. Hofstraat, and
4:Unit 4.1. G.J. Brakenhoff. 2004. Image calibration
Schaufele, F., I. Demarco, and R.N. Day. 2005. in fluorescence microscopy. J. Microsc.
FRET imaging in the widefield microscope. 216:15–24.
In Molecular Imaging: FRET Microscopy
and Spectroscopy. A. Periasamy and R.N.
Day, editors. Oxford University Press.
72–94.
Sekar, R.B., and A. Periasamy. 2003. Fluorescence
resonance energy transfer (FRET) micros-
copy imaging of live cell protein localiza-
tions. J. Cell Biol. 160:629–633.
Shaner, N.C., P.A. Steinbach, and R.Y. Tsien. 2005.
A guide to choosing fluorescent proteins.
Nat. Methods. 2:905–909.
Siegel, R.M., F.K. Chan, D.A. Zacharias, R.
Swofford, K.L. Holmes, R.Y. Tsien, and
M.J. Lenardo. 2000. Measurement of mo-
lecular interactions in living cells by fluor­
escence resonance energy transfer between
variants of the green fluorescent protein. Sci.
STKE. 2000:PL1.
Snapp, E.L., N. Altan, and J. Lippincott-Schwartz.
2003. Measuring protein mobility by photo-
bleaching GFP chimeras in living cells. Curr.
Protoc. Cell Biol. Chapter 21:Unit 21.1.
Sprague, B.L., and J.G. McNally. 2005. FRAP analy­

Downloaded from jcb.rupress.org on June 30, 2009


sis of binding: proper and fitting. Trends
Cell Biol. 15:84–91.
Spring, K.R. 2007. Cameras for digital microscopy.
Methods Cell Biol. 81:171–186.
Straight, A.F. 2007. Fluorescence protein appli-
cations in microscopy. Methods Cell Biol.
81:93–113.
Stryer, L. 1978. Fluorescence energy transfer as a
spectroscopic ruler. Annu. Rev. Biochem.
47:819–846.
Swedlow, J.R., K. Hu, P.D. Andrews, D.S. Roos,
and J.M. Murray. 2002. Measuring tubulin
content in Toxoplasma gondii: a comparison
of laser-scanning confocal and wide-field
fluorescence microscopy. Proc. Natl. Acad.
Sci. USA. 99:2014–2019.
Tommre, D., and J.B. Pawley. 2006. Disk-scanning
confocal microscopy. In Handbook of Bio­
logical Confocal Microscopy. J.B. Pawley,
editor. Springer-Verlag New York, Inc., New
York. 221–237.
Tsien, R.Y., L. Ernst, and A. Waggoner. 2006.
Fluorophores for confocal microscopy: pho-
tophysics and photochemistry. In Handbook
of Biological Confocal Microscopy. J.B.
Pawley, editor. Springer-Verlag New York,
Inc., New York. 338–348.
Vogel, S.S., C. Thaler, and S.V. Koushik. 2006.
Fanciful FRET. Sci. STKE. 2006:re2.
Wallace, W., L.H. Schaefer, and J.R. Swedlow.
2001. A workingperson’s guide to deconvo-
lution in light microscopy. Biotechniques.
31:1076–1097.
Waters, J.C. 2007. Live-cell fluorescence imaging.
Methods Cell Biol. 81:115–140.
Waters, J.C., and J.R. Swedlow. 2007. Interpreting
fluorescence microscopy images and mea-
surements. In Evaluating Techniques in
Biochemical Research. D. Zuk, editor. Cell
Press, Cambridge, MA. 37–42.
Wolf, D.E., C. Samarasekera, and J.R. Swedlow.
2007. Quantitative analysis of digital
microscope images. Methods Cell Biol.
81:365–396.
Wu, J.Q., and T.D. Pollard. 2005. Counting cyto­
kinesis proteins globally and locally in fission
yeast. Science. 310:310–314.
Yildiz, A., and P.R. Selvin. 2005. Fluorescence
imaging with one nanometer accuracy:

 1148 JCB • VOLUME 185 • NUMBER 7 • 2009

Вам также может понравиться