Chapter 11 – Oocyte Retrieval and Embryo Culture




Abstract




Every individual treatment cycle involves a number of different stages and manipulations in the laboratory, and each case must be assessed and prepared for in advance; the afternoon prior to the procedure (the day after hCG administration) is a convenient time to make the preparations. The laboratory staff should ensure that all appropriate consent forms have been signed by both partners, including consent for special procedures and storage of cryopreserved embryos. Details of any previous assisted conception treatment should be studied, including response to stimulation, number and quality of oocytes, timing of insemination, fertilization rate, embryo quality and embryo transfer procedure, and judgments regarding whether any parameters at any stage could be altered or improved in the present cycle can be assessed. The risk of introducing any infection into the laboratory via gametes and samples must be absolutely minimized: screening tests such as human immunodeficiency virus (HIV 1 and 2: Anti-HIV 1, 2) and hepatitis B (HbsAg/Anti-HBc) and C (Anti-HCV-Ab) should be confirmed, as well as any other tests indicated by the patients’ history (e.g., HTLV-I antibody, RhD, malaria, Trypanosoma cruzi, Zika virus). If donor gametes are to be used, additional tests for the donor are required: chlamydia, cytomegalovirus and a validated testing algorithm to exclude the presence of active infection with Treponema pallidum for syphilis testing.





Chapter 11 Oocyte Retrieval and Embryo Culture




Preparation for Each Case


Every individual treatment cycle involves a number of different stages and manipulations in the laboratory, and each case must be assessed and prepared for in advance; the afternoon prior to the procedure (the day after hCG administration) is a convenient time to make the preparations. The laboratory staff should ensure that all appropriate consent forms have been signed by both partners, including consent for special procedures and storage of cryopreserved embryos. Details of any previous assisted conception treatment should be studied, including response to stimulation, number and quality of oocytes, timing of insemination, fertilization rate, embryo quality and embryo transfer procedure, and judgments regarding whether any parameters at any stage could be altered or improved in the present cycle can be assessed. The risk of introducing any infection into the laboratory via gametes and samples must be absolutely minimized: screening tests such as human immunodeficiency virus (HIV 1 and 2: Anti-HIV 1, 2) and hepatitis B (HbsAg/Anti-HBc) and C (Anti-HCV-Ab) should be confirmed, as well as any other tests indicated by the patients’ history (e.g., HTLV-I antibody, RhD, malaria, Trypanosoma cruzi, Zika virus). If donor gametes are to be used, additional tests for the donor are required: chlamydia, cytomegalovirus and a validated testing algorithm to exclude the presence of active infection with Treponema pallidum for syphilis testing.




Laboratory Preparation: Checklist for Each Case




  • Results of viral screening tests for both partners



  • Consent forms signed by both partners



  • Specific details or instructions regarding insemination, cryopreservation, number of embryos for transfer, etc.



  • Results of semen assessments; note any special features or precautions for semen collection or preparation



  • Previous history, current response to stimulation; note any special features of ultrasound scans, endocrine assays or IVF laboratory results



  • Current cycle history: number of follicles, endocrine parameters, potential ovarian hyperstimulation syndrome (OHSS)


Laboratory case notes, media, culture vessels and tubes for sperm preparation, with clear and adequate labeling throughout, are prepared in advance of each case. All labeling should have a minimum of the patient’s full name and a unique identifier, for example, patient number. When donor sperm is used, the donor code must uniquely identify that specific donor. Tissue culture dishes or plates can be equilibrated in the culture incubator overnight. The choice of culture system used is a matter of individual preference and previous experience; microdroplets under oil, four-well dishes and organ culture dishes are amongst those most commonly used.



Culture Dish Preparation: General Considerations



Osmolality

In vitro, gametes and embryos are very sensitive to even small external increases in the osmolality of their environment, with potentially dramatic effects on viability. Inappropriate handling of the culture system can lead to changes in osmolality that are sufficient to jeopardize embryonic development, and osmolality of a culture system is influenced by the methods used in dish preparation. Minimizing the risk of evaporation is crucial when preparing culture dishes: time, temperature, air flow, volume of media and oil overlay, as well as the method of preparation, must all be considered in order to avoid osmolality increases that may jeopardize embryo development (see Swain et al., 2012; Elder et al., 2015, Chapter 4). Evaporation can be minimized by taking culture media directly from the refrigerator immediately prior to dish preparation, so that the droplets are prepared with cold medium. In open cultures (without oil), a humidified atmosphere in the incubator is essential to prevent evaporation and increase in osmolality.


A number of systems are now available that differ from the traditional methods of placing droplets onto flat dish surfaces, such as microwell dishes with indents manufactured into the surface, GPS Corral dishes and dishes designed for use with time-lapse systems and incubators. Osmolality is an important factor that will influence the successful application of these systems. The depth of medium, surface area exposed to the oil overlay and depth of the overlay may all influence the osmolality of the culture medium. Preparation of dishes for time-lapse imaging requires special care in order to make sure that no bubbles are present. Devices used for microfluidic culture are particularly vulnerable to detrimental osmolality shifts due to evaporation; thin hybrid polydimethylsiloxane (PDMS)-Parylene membranes are recommended to circumvent this problem (Heo et al., 2007).



Practical Tips for Keeping Osmolality Constant during Preparation of Dishes



  1. 1. Take note of the osmolality and osmolyte content of the medium to be used; osmolytes act to maintain fluid balance within oocytes/blastomeres, and some media may have a lower concentration of osmolytes than others (HTF has no osmolytes). If there is a risk of increase in osmolality during preparation, consider using a medium that has a lower osmolality.



  2. 2. Work at room temperature, using medium that has been taken out of the refrigerator immediately prior to dish preparation.



  3. 3. Consider droplet size. Using larger volumes of media in the culture system will minimize the chance of large shifts in osmolality; droplet sizes less than 20 µl are not recommended.



  4. 4. Carefully note the date of first opening the bottle of medium in relation to the date of dish preparation: osmolality can change if the bottle has been opened repeatedly.



  5. 5. Note the time/date of dish preparation: the likelihood of evaporation with resulting rise in osmolality is higher in dishes that have been in the incubator for long periods.



  6. 6. Prepare droplets as quickly as possible, reducing the number of droplets per dish if necessary.



  7. 7. If droplets are made before adding the oil overlay, replace original droplets with droplets of fresh medium in exactly the same spot after overlaying with oil.



  8. 8. If sufficient staff are available, the dishes can be prepared by two technicians: one to pipette the droplets into the dish and the second to immediately cover the droplets with oil. Ensure that good laboratory practice and sterility is rigorously maintained throughout.



  9. 9. Be aware that osmolality will change more rapidly with ‘open’ culture and without humidified incubators.



  10. 10. Reduce airflow during dish preparation by turning the laminar air flow off. However, this is recommended only if background air has a suitable grade of purity. Since it is essential that sterility is maintained during dish preparation, an alternative to consider is reducing the air flow velocity by setting it to the lowest level during dish preparation.



  11. 11. Use washed/humidified oil: in general, there is always an equilibrium between the components of two liquid phases. Unless the oil is first washed with the same medium, an oil interface with albumin-carrying lipids in the medium may extract all lipophilic elements present in the medium. Washing can be carried out by gently shaking oil with medium in a ratio of 1(oil):2(medium) and leaving the mixture to settle overnight, until the two phases are clearly separate.



Microdroplets under Oil




  • Pour previously equilibrated mineral oil into 60-mm Petri dishes that are clearly marked with each patient’s full name.



  • Using either a Pasteur pipette or adjustable pipettor and sterile tips, carefully place eight or nine droplets of medium around the edge of the dish. One or two droplets may be placed centrally, to be used as wash drops (alternatively, droplets can be placed first, and then overlaid with oil – see above).



  • Examine the follicular growth records to assess approximately how many drops/dishes should be prepared; each drop may contain one or two oocytes. Droplet size can range from 50 to 250 µL per droplet.



Four-Well Plates


This system may also be used in combination with an overlay of equilibrated mineral oil.




  • Prepare labeled and numbered plates containing 0.5–1 mL of tissue culture medium; each well is normally used to incubate up to three oocytes.



  • Equilibrate overnight: If used without an oil overlay, the incubator must be humidified.



  • Small Petri dishes with approximately 2 mL of HEPES-containing medium may also be prepared, to be used for washing oocytes immediately after identification in the follicular aspirates (media containing HEPES should not be equilibrated in a CO2 atmosphere).


Organ culture (center well) dishes can also be used for group culture: place three 250-µL drops in the center well, and three to four oocytes in each drop.



New Directions for Culture Systems


The quest for defining better media and culture systems has traveled from simple salt solutions in gas-filled desiccators to computer-controlled microchambers with continuous renewal of media (Figure 11.1). Microfluidic systems extend the strategy behind sequential media by aiming to mimic dynamic changes in the tubal environment of the embryo during its transit toward the uterus (Swain et al., 2012, 2013; da Rocha & Smith, 2012, 2014). At present, the use of such chambers is limited by high costs, and by problems associated with maintaining sterile conditions in pumps and tubing. Dynamic culture platforms that involve tilting and vibrating (‘rock and roll’) have also been proposed as being of potential benefit, by disrupting media gradients that form around cells or stimulating mechano-sensitive signaling pathways that might stimulate embryo growth (Swain, 2013).





Figure 11.1 Example of a microfluidic device used for embryo culture. 1: Reservoir for fresh medium; 2: pneumatic microchannel; 3: cell culture chamber; 4: waste medium reservoir; 5: normally closed valve; 6: airtube insertion hole; 7: microchannel for medium; 8: pneumatically driven membrane-based micropump.


Adapted from da Rocha & Smith (2012), with kind permission from Springer Science+Business Media.



Laboratory Supplies for Labeling and Preparation




  • Pots for semen collection



  • Test-tubes for sperm preparation: conical tubes, small and large round-bottomed tubes



  • Aliquots of media for each patient



  • Culture vessels for overnight equilibration



  • Paperwork for recording case details and results



Oocyte Retrieval (OCR) and Identification


A Class II biological safety cabinet is recommended for handling of follicular aspirates to avoid risk of infection, but be aware that the laminar flow of air can have a dramatic cooling effect on the samples.


Prior to the follicle aspiration procedure:




  1. 1. Ensure: heating blocks, stages and trays are warmed to a temperature that will maintain the medium in the dishes at 37°C; media to be used for flushing/rinsing must also be warmed and equilibrated to correct pH.



  2. 2. Prewarm collection test-tubes and 60-mm Petri dishes for scanning aspirates.



  3. 3. Prepare a sterile Pasteur pipette plus holder, a fine-drawn blunt Pasteur pipette as a probe for manipulations, and 1-mL syringes with attached needles for dissection.



  4. 4. Check patient ID, confirm names and unique identifiers on dishes and laboratory case notes with medical notes.


If follicular aspirates cannot be examined immediately, they should be collected into test-tubes that are completely filled with fluid, tightly sealed and rigorously maintained at 37°C until they reach the laboratory. Aliquot the contents of each test-tube into two or three Petri dishes, forming a thin layer of fluid that can be quickly, carefully and easily scanned for the presence of an oocyte, using a stereo dissecting microscope with transmitted illumination base and heated stage. Low-power magnification (×6–12) can be used for scanning the fluid, and oocyte identification verified using higher magnification (×25–50). Always work quickly and carefully, with rigid attention to sterile technique, maintaining correct temperature and pH at all times.



Oocyte Identification


The oocyte usually appears within varying quantities of cumulus cells and, if very mature, may be pale and difficult to see. (Immature oocytes are dark and also difficult to see.) Granulosa cells are clearer and more ‘fluffy,’ present in amorphous, often iridescent clumps. Blood clots, especially from the collection needle, should be carefully dissected with 23-gauge needles to check for the presence of cumulus cells.


The presence of blood clots within the cumulus–oocyte complex (COC) may be a reflection of poor follicular development, with an effect on the competence of the corresponding oocyte (Ebner, 2008). When a COC is identified, assess its stage of maturity by noting the volume, density and condition of the surrounding cumulus cells and the expansion of coronal cells. It is unlikely that the oocyte itself can be seen, since it will most commonly be surrounded by cumulus cells. However, when an oocyte can be observed with minimal cumulus cells, the presence of a single polar body indicates that it has reached the stage of metaphase II. The appearance of the COC can be used to classify the oocyte according to the following scheme (Figure 11.2):




  1. 1. Germinal vesicle: the oocyte is very immature. There is no expansion of the surrounding cells, which are tightly packed around the oocyte. A large nucleus (the germinal vesicle) is still present and may occasionally be seen with the help of an inverted microscope. Maturation may occasionally take place in vitro from this stage, and the COC can be assessed later in the day prior to insemination (Figure 11.2e).



  2. 2. Metaphase I: the oocyte is surrounded by a tightly apposed layer of corona cells; tightly packed cumulus with little extracellular matrix may surround this with a maximum size of approximately five oocyte diameters. If the oocyte can be seen, it no longer shows a germinal vesicle. The absence of a polar body indicates that the oocyte is in metaphase I, and these immature oocytes may be preincubated for 6–24 hours before insemination (Figure 11.2f).



  3. 3. Metaphase II:




    1. a. Preovulatory (harvested from Graafian follicles): this is the optimal level of maturity, appropriate for successful fertilization. Coronal cells are still apposed to the oocyte but are fully radiating; one polar body has been extruded. The cumulus has expanded into a fluffy viscous mass that can be easily stretched, with abundant extracellular matrix (Figure 11.2a and b).



    2. b. Mature: the oocyte can often be seen clearly as a pale orb; little coronal material is present and is dissociated from the oocyte. The cumulus is very profuse but is still cellular. The latest events of this stage involve a condensation of cumulus into small black (refractile) drops, as if a tight corona is reforming around the oocyte. The perivitelline space often shows granularity (Figure 11.2h).



    3. c. Luteinized: the oocyte is very pale and often is difficult to find. The cumulus has broken down and becomes a gelatinous mass around the oocyte. These oocytes have a low probability of fertilization, and are usually inseminated with little delay (Figure 11.2).



    4. d. Atretic: the oocyte is very dark and can be difficult to identify. Granulosa cells are fragmented and have a lace-like appearance.







(a) Cumulus–oocyte complex visualized under dissecting microscope, ×25 magnification.





(b) Phase-contrast image of mature metaphase II cumulus–oocyte complex, first polar body visible.





(c) Cumulus–oocyte complex with clumped refractile areas indicating signs of luteinization.





(d) Empty zona pellucida. Phase-contrast images after denudation.





(e) Germinal vesicle.





(f) Metaphase I oocyte, polar body not extruded.





(g) Preovulatory metaphase II oocyte, polar body extruded.





(h) Postmature oocyte, showing granularity in the perivitelline space.





(i) Dysmorphic metaphase II oocyte showing a large necrotic first polar body.


Images (e), (f), (g) courtesy of Julia Uraji; Images (h) and (i) courtesy of Thomas Ebner, Austria.


Figure 11.2 Images after oocyte retrieval, before and after hyaluronidase denudation. Cumulus–oocyte complexes, without denudation.


Gross morphological assessment of oocyte maturity is highly subjective and open to inaccuracies. In preparation for intracytoplasmic sperm injection (ICSI), the oocytes are completely denuded of surrounding cells using hyaluronidase, allowing accurate assessment of nuclear maturity and cytoplasm; this process has made it apparent that gross COC morphology does not necessarily correlate with nuclear maturity, and there is considerable conflict in the data available regarding the association between oocyte morphology and treatment outcome (see Ebner, 2006, for review). A number of dysmorphic features can be identified in denuded oocytes, including areas of necrosis, organelle clustering, vacuolation or accumulating aggregates of smooth endoplasmic reticulum (sER). Anomalies of the zona pellucida and nonspherical oocytes can also be seen. In practice, a wide variety of unusual and surprising dysmorphisms are often observed. A collection of interesting photographs, together with patient clinical details and histories, can be seen in the Atlas of Oocytes, Zygotes and Embryos in Reproductive Medicine by Van den Bergh et al. (2012).


Some features of dysmorphism may be associated with the endocrine environment during ovarian stimulation, in particular the structure of the zona pellucida and/or oolemma (Ebner, 2002, 2006). Although aberrations in the morphology of oocytes are not necessarily of any consequence to fertilization or early cleavage after ICSI, it is possible that embryos generated from dysmorphic oocytes have a reduced potential for implantation and further development. Repeated appearance of some dysmorphic features such as sER aggregation, central granulation or vacuoles in an individual patient’s oocyte cohort may indicate an underlying intrinsic problem in the process of oocyte development within the ovary.




Dysmorphic Oocyte Features


See Figure 11.3.




  • Irregular shape



  • Areas of necrosis in the cytoplasm



  • Cytoplasmic granularity



  • Organelle clustering



  • Aggregates of sER



  • Vacuoles/vesicles



  • Anomalies of the zona pellucida





Figure 11.3 Normal and dysmorphic oocytes (A, B). Normal appearing oocytes with no visually outstanding features (C–F). Varying degrees of organelle clusters (*) (central granularity) observed from mild to very severe. (G, N) Aggregation (arrows) of smooth endoplasmic reticulum as a flat, clear disc in the middle of the cytoplasm of the oocyte. (H) A dark ‘horse-shoe-shaped’ (large arrow) cytoplasmic inclusion. (I, J, K) Varying degrees (mild to severe) of fluid-filled vacuoles within the cytoplasm. (L) Organelle cluster with fragmented polar body (arrow) and increased perivitelline debris (*) and space. (K–M) Combination of cytoplasmic dysmorphisms and extracytoplasmic phenotypes. PB1 = first polar body.


From Meriano et al. (2001), with permission from Oxford University Press.

Van Blerkom and Henry (1992) reported aneuploidy in 50% of oocytes with cytoplasmic dysmorphism; it is not clear whether oocyte aneuploidy is a fundamental developmental phenomenon or a patient-specific response to induced ovarian stimulation. In 1996, the same group related the oxygen content of human follicular fluid to oocyte quality and subsequent implantation potential. They propose that low oxygen tension associated with poor blood flow to follicles lowers the pH and produces anomalies in chromosomal organization and microtubule assembly, which might cause segregation disorders. Measurement of blood flow to individual follicles by power color Doppler ultrasound (Gregory and Leese, 1996) confirmed the observations of Van Blerkom et al. (1997) in correlating follicular blood flow with implantation; the incidence of triploid zygotes was also found to be significantly higher when oocytes were derived from follicles with poor vascularity. Follicular vascularity may also influence free cortisol levels in follicular fluid by promoting its diffusion across the follicle boundary. However, a robust model that can be applied for clinical assessment is not yet available (Mercé et al., 2006).



Insemination


Oocytes are routinely inseminated with a concentration of 100 000 progressively motile sperm per milliliter. If the prepared sperm show suboptimal parameters of motility or morphology, the insemination concentration may be accordingly increased. Some reports have suggested that the use of a high insemination concentration of up to 300 000 progressively motile sperm per milliliter may be a useful prelude before deciding upon ICSI treatment for male factor patients. Traditionally, inseminated oocytes were incubated overnight in the presence of the prepared sperm sample; however, sperm binding to the zona pellucida normally takes place within 1–3 hours of insemination, and fertilization occurs very rapidly thereafter. A few hours of sperm–oocyte contact yields the same time course of events that is observed after overnight incubation, and oocytes can be washed free of excess sperm after 3 hours’ incubation (Gianaroli et al., 1996; Ménézo and Barak, 2000).


For a culture system of microdroplets under oil, each oocyte is transferred into a drop containing motile sperm at a concentration of approximately 100 000 sperm/mL. In a four-well system, a measured volume of prepared sperm is added to each well, to a final concentration of approximately 100 000 progressively motile sperm per well.




Insemination


For microdroplets under oil, the oil overlays for insemination dishes must be prepared earlier, so that there is at least 4–6 hours of equilibration time.




  1. 1. Prepare a dilution of prepared sperm, containing 100 000 motile sperm/mL




    • Assess a drop of the dilution on a glass slide, under ×10 magnification; at least 20 motile sperm should be visible in the field.



    • Equilibrate the suspension at 37°C for 30 minutes, 5% CO2.



    • Place droplets of the sperm suspension under the previously prepared and equilibrated oil overlays.



    • Examine each oocyte before transfer to the insemination drop, and dissect the cumulus to remove bubbles, large clumps of granulosa cells or blood clots if necessary.




  2. 2. If oocytes are in premeasured culture droplets, e.g., 240 μL, add 10 μL of a prepared sperm suspension that has been adjusted to 2.5 × 106/mL.




    • Final concentration = approximately 100 000 sperm/mL, or 25 000 sperm per oocyte.




  3. 3. Prepare labeled 35-mm Petri dishes containing equilibrated oil, to be used for culture of the zygotes after scoring for fertilization the following day.


Four-well dishes and organ culture dishes




  • Add 0.5–1.0 mL of prepared sperm suspension to each well or drop.



  • Total: approximately 100 000 motile sperm per well/drop.



Scoring of Fertilization on Day 1



Dissecting Fertilized Oocytes


Inseminated oocytes are dissected 17–20 hours following insemination in order to assess fertilization. Oocytes at this time are normally covered with a layer of dispersed coronal and cumulus cells, which must be carefully removed so that the cell cytoplasm can be examined for the presence of two pronuclei and two polar bodies, indicating normal fertilization. The choice of dissection procedure is a matter of individual preference, and sometimes a combination of methods may be necessary for particular cases. Whatever the method used, it must be carried out carefully, delicately and speedily, taking care not to expose the fertilized oocytes to changes in temperature and pH. Scoring for pronuclei should be carried out within the appropriate time span, before pronuclei merge during syngamy: cleaved embryos with abnormal fertilization are indistinguishable from those with two pronuclei.



Dissection Techniques





  1. 1. Narrow-gauge pipetting: narrow-gauge pipettes can be made (see below), but commercial hand-held pipetting devices are simpler and more convenient. ‘Flexipet’ and ‘Stripper’ are hand-held pipetting devices for cumulus/corona removal, with sterile disposable polycarbonate capillaries of specified inner diameters ranging from 135 up to 175 or 600 µm. Variations that incorporate a capillary that attaches to a tiny pressure ‘bulb’ inserted into a hollow metal tube are also available.




    • Use the microscope at ×25 magnification, and choose a tip with a diameter slightly larger than the oocyte. (A tip that is too small will damage the oocyte; therefore, take care in selecting the appropriate diameter.)



    • Aspirate approximately 2 cm of clean culture medium into the tip, providing a protective buffer. This allows easy flushing of the oocyte and prevents it from sticking to the inside surface of the tip.



    • Place the tip over the oocyte and gently aspirate it into the shaft.



    • If the oocyte does not easily enter, change to a larger diameter pipette. (However, if the diameter is too large, it will be ineffective for cumulus removal.)



    • Gently aspirate and expel the oocyte through the pipette, retaining the initial buffer volume, until sufficient cumulus and corona is removed to allow clear visualization of the cell cytoplasm and pronuclei.




  2. 2. Needle dissection: use two 26-gauge needles attached to 1-mL syringes, microscope at ×25 magnification. Use one needle as a guide, anchoring a piece of cellular debris if possible; slide the other needle down the first one, ‘shaving’ cells from around the zona pellucida, with a scissors-like action.



Before commercial hand-held denudation devices were available, a technique known as ‘Rolling’ was sometimes used, and a description of this method is included here for historical interest only.




  • Use one 23-gauge needle attached to a syringe, and a fine glass probe.



  • With the microscope at ×12 magnification, use the needle to score lines in each droplet on the base of the plastic dish.



  • Adjust the magnification to ×25, and push the oocyte gently over the scratches with a fire-polished glass probe until the adhering cells are teased away. This technique was useful in removing adherent sticky blood clots.


Great care must be taken with any technique to avoid damaging the zona pellucida or the oocyte either by puncture or overdistortion. Breaks or cracks in the zona can sometimes be seen, and a small portion of the oocyte may extrude through the crack. (This may have occurred during dissection or during the aspiration process.) Occasionally the zona is very fragile, fracturing or distorting at the slightest touch; it is probably best not to continue the dissection in these cases.




Making Narrow-Gauge Pipettes


The preparation of finely drawn pipettes with an inner diameter slightly larger than the circumference of an oocyte is an acquired skill which requires practice and patience.




  • Hold both ends of the pipette, and roll an area approximately 2.5 cm below the tapered section of the pipette over a gentle flame (Bunsen or spirit burner).



  • As the glass begins to melt, quickly pull the pipette in both directions to separate.



  • Before the glass has a chance to cool, carefully and quickly break the pipette at an appropriate position.



  • The tip must have a clean break, without rough or uneven edges; these will damage the oocyte during dissection.



  • Examine the tip of each pipette to ensure that it is of accurate diameter, with smooth clean edges.



Pronuclear Scoring


An inverted microscope is recommended for accurate scoring of fertilization; although the pronuclei can be seen with dissecting microscopes, it can often be difficult to distinguish normal pronuclei from vacuoles or other irregularities in the cytoplasm. Normally fertilized oocytes should have two pronuclei, two polar bodies, regular shape with intact zona pellucida and a clear healthy cytoplasm. A variety of different features may be observed: the cytoplasm of normally fertilized oocytes is usually slightly granular, whereas the cytoplasm of unfertilized oocytes tends to be completely clear and featureless. The cytoplasm can vary from slightly granular and healthy-looking, to brown or dark and degenerate. The shape of the oocyte may also vary, from perfectly spherical to irregular (see Figure 11.3). A clear halo of peripheral cytoplasm 5–10 mm thick is an indication of good activation and reinitiation of meiosis. The pattern and alignment of nucleoli has also been thought to be significant (Scott and Smith, 1998; Tesarik and Greco, 1999).


Approximately 5% of fertilized oocytes in human IVF routinely show abnormal fertilization, with three or more pronuclei visible; this is attributed to polyspermy, or nonextrusion of the second polar body. Fluorescent in-situ hybridization (FISH) analysis indicates that 80–90% of these zygotes are mosaic after cleavage. Single pronucleate zygotes obtained after conventional IVF analyzed by FISH to determine their ploidy reveal that a proportion of these zygotes are diploid (Levron et al., 1995). It seems that during the course of their interaction, it is possible for human gamete nuclei to associate together and form diploid, single pronucleate zygotes. These findings may indicate a variation of human pronuclear interaction during syngamy, and the authors suggest that single pronucleate zygotes which develop with normal cleavage may be selected for transfer in cases where no other suitable embryos are available.


Details of morphology and fertilization should be recorded for each zygote, for reference when choosing embryos for transfer. Remove zygotes with normal fertilization at the time of scoring from the insemination drops or wells, transfer into new dishes or plates containing pre-equilibrated culture medium, and return them to the incubator for a further 24 hours of culture. Those with abnormal fertilization such as multipronucleate zygotes should be discarded, so that there is no possibility of their being selected for embryo transfer; after cleavage, these are indistinguishable from normally fertilized oocytes.


Although the presence of two pronuclei confirms fertilization, their absence does not necessarily indicate fertilization failure, and may instead represent either parthenogenetic activation or a delay in timing of one or more of the events involved in fertilization (Figure 11.4). Numerous studies have accumulated evidence to demonstrate that up to 40% of oocytes with no sign of fertilization 17–27 hours after insemination may have the appearance of morphologically normal embryos on the following day, with morphology and cleavage rate similar to that of zygotes with obvious pronuclei on Day 1. However, around a third of these zygotes may subsequently arrest on Day 2 (Plachot et al., 1993). Cytogenetic analysis of these embryos reveals a higher incidence of chromosomal anomalies and a high rate of haploidy, confirming parthenogenetic activation (Plachot et al., 1988, 1993).






(a) Normal fertilization: two pronuclei, two polar bodies.





(b) Abnormal fertilization: three pronuclei.





(c) Abnormal fertilization, no pronuclei, two polar bodies.





(d) Zygote showing two pronuclei, numerous vacuoles and irregular perivitelline space, illustrating that severely dysmorphic oocytes are capable of fertilization.


With thanks to Marc van den Bergh.


Figure 11.4 Phase-contrast micrographs of fertilized human oocytes.


Delayed fertilization with the appearance of pronuclei on Day 2 may also be observed, and these embryos also tend to have an impaired developmental potential. Delayed fertilization can be attributed to morphological or endocrine oocyte defects in some cases, and to sperm defects in others. No obvious association with either oocyte or sperm defects can be found in a number of cases (Oehninger et al., 1989).



Reinsemination


Reinsemination of oocytes that fail to demonstrate clear pronuclei at the time of scoring for fertilization is a practice that has been widely questioned scientifically. Fertilization or cleavage may subsequently be observed on Day 2, but this may be as a consequence of the initial insemination, and the delay in fertilization may be attributed either to functional disorders of the sperm, or maturation delay of the oocyte. These embryos generally have a poor prognosis for implantation.



‘Rescue’ ICSI


In cases of total failure of fertilization after IVF or ICSI, ‘rescue’ is sometimes attempted as a last resort by injecting the oocytes on Day 1. This practice is banned in some countries such as the United Kingdom, since it cannot be certain if a sperm has already entered the oocyte and fertilization is delayed. Others reserve rescue ICSI only for cases where there is complete failure to fertilize following conventional IVF. However, it is now clear that successful embryo development is crucially dependent upon events surrounding the timing of fertilization, as well as on cytoplasmic maturity: extended culture risks numerous negative effects on the oocyte. Although fertilization can sometimes be achieved via ‘rescue ICSI,’ developmental potential of the embryos is very poor, with minimal chance of pregnancy (see Beck-Fruchter et al., 2014, for review).



Selection of Pronucleate Embryos for Cryopreservation


Legislation in some countries forbids embryo freezing but allows cryopreservation at the zygote stage, before syngamy. Zygotes to be frozen should have a regular outline, distinct zona and clearly visible pronuclei. The cryopreservation procedure must be initiated while the pronuclei are still visible, before the onset of syngamy (see Chapter 12).



Selection of Embryos for Transfer


Historically, embryo transfer was carried out 2 days (approximately 48–54 hours) after oocyte retrieval, but transfer has been carried out from as early as 1 hour post-ICSI (AOT, activated oocyte transfer, Dale et al., 1999) to 5 days later, at the blastocyst stage. Trials of zygote transfer on Day 1 also achieved acceptable pregnancy rates (Scott and Smith, 1998; Tesarik and Greco, 1999; Tesarik et al., 2000); it seems that the specific timing of transfer may not be crucial for the human implantation process. On Day 2, cleaved embryos may contain from two to six blastomeres. Embryo transfer 1 day later, on Day 3, or on Day 5 at the blastocyst stage is advocated as a means of selecting embryos with higher implantation potential, by the elimination of those that arrest at earlier cleavage stages in vitro.


Two major problems continue to hinder the effectiveness of ART treatment: low implantation rates and a high incidence of multiple pregnancies. Poor endometrial receptivity and adverse uterine contractions can both contribute to early embryo loss, but the low efficiency of assisted conception is widely attributed to genetic defects in the embryo. More than 40% of ART-derived embryos are known to harbor chromosomal abnormalities. Errors in meiotic and mitotic segregation of chromosomes in the oocyte and during the cleavage of early embryos can lead to different patterns of aneuploidy, including polyploidy and chaotic mosaics, which account for around one-third of aneuploidies involving more than one chromosome per cell. However, despite the fact that grossly abnormal chromosome complements are lethal, in most cases the morphology of embryos that are genetically normal does not differ markedly from those with aneuploid, polyploid or mosaic chromosomal complements. Consequently, genetically abnormal embryos after IVF or ICSI may be graded as suitable for transfer using subjective selection criteria. Developing a reliable diagnostic test that can be used to identify embryos with the greatest developmental competence continues to be a major priority in human ART, in the hope of eventually selecting a single embryo that is likely to result in a healthy live birth following transfer.


In selecting embryos for transfer, the limitations of evaluating embryos based on morphological criteria alone are well recognized: correlations between gross morphology and implantation are weak and inaccurate, unless the embryos are clearly degenerating/fragmented. Objective criteria for evaluating embryos are available in laboratories with research facilities, but may be out of reach for a routine clinical IVF laboratory without access to specialized equipment and facilities. Objective measurements of human embryo viability that have been applied historically include:




  • High-resolution videocinematography



  • Computer-assisted morphometric analysis



  • Blastomere or polar body biopsy for cytogenetic analysis



  • Culture of cumulus cells



  • Oxygen levels in follicular fluid/perifollicular vascularization



  • Distribution of mitochondria and ATP levels in blastomeres



  • Molecular approaches:




    • Metabolic assessment of culture media (amino acid profiling, metabolomics)



    • Gene expression/expression of messenger RNA (mRNA) in cumulus cells and/or embryos.





Sequence of Events Observed by Time-Lapse Cinematography


From Mio and Maeda (2008).




  • Day 0 = day of OCR, insemination approximately 4–5 hours post OCR.



Activation of the zygote genome begins at this stage, with massive increase in transcription and translation; this cell cycle requires a full 24 hours.


By 96 hours post-insemination, on Day 4: compacting morula stage (around 32 cells).


Day 5: differentiation to blastocyst stage. Cell number may vary considerably, from 50 to 100–120 cells for early expanding blastocysts.


Hatching may start as early as the morning of Day 5 but is usually observed on Day 6/7.



High-Resolution Videocinematography


As early as 1989, Cohen et al. carried out a classic experiment that aimed to clearly define morphological criteria that might be used for embryo assessment, using a detailed analysis of videotaped images. Immediately before embryo transfer, embryos were recorded on VHS for 30–90 seconds, at several focal points, using Nomarski optics and an overall magnification of ×1400. The recordings were subsequently analyzed by observers who were unaware of the outcome of the IVF procedure, and they objectively assessed a total of 11 different parameters:




























Objective Parameters Assessed on Videocinematography
Cell organelles visible Cellular extrusions
Blastomeres all intact Cytoplasmic vacuoles
Identical blastomere size Blastomeres contracted
Smooth membranes % variation in zona thickness
Dark blastomeres % extracellular fragments
Cell–cell adherence

Nine parameters were judged (+) or (–), and variation in zona thickness and percentage of extracellular fragments were given a numerical value. Analysis of these criteria showed no clear correlation with any intracellular features of morphology, but that the most important predictor of fresh embryo implantation was the percentage of variation in thickness of the zona pellucida. Embryos with a thick, even zona had a poor prognosis for implantation; those whose zona had thin patches also had ‘swollen,’ more refractile blastomeres, and had few or no fragments. This observation was one of the parameters that led the group to introduce the use of assisted hatching (see Chapter 13). In analyzing frozen-thawed embryos, the best predictor of implantation was cell–cell adherence. The proportion of thawed embryos with more than one abnormality (77%) was higher than that of fresh embryos (38%) despite similar implantation rates (18% versus 15%).


Nearly 30 years have elapsed since these observations were published, and the quest to identify specific morphological markers of embryo implantation potential still continues – now with the help of more sophisticated technology to measure both properties of the zona and detailed embryo morphology.



Zona Pellucida Birefringence


Polarized microscopy allows three layers to be distinguished in the zona, with the innermost layer showing the greatest birefringence (i.e., a higher level of light retardance). Several studies have investigated a possible correlation between this zona property and implantation potential (Montag et al., 2008; Madaschi et al., 2009); there is no doubt that properties of the zona may be important in assessing oocyte/embryo potential, particularly in response to exogenous FSH stimulation, but further studies are required in order to establish a clear correlation.



Computer-Assisted Morphometric Analysis


High-resolution digital images of embryos can be assessed in detail with the help of computer-assisted multilevel analysis, which provides a three-dimensional picture of embryo morphology. A system with a computer-controlled motorized stepper mounted on the microscope will automatically focus through different focal planes in the embryo to produce a sequence of digital images. Automatic calculations of morphometric information from the image sequences describe features and measurements of each embryo, including size of nucleus and blastomeres and their spatial positions within the embryo, as well as features of the zona pellucida; all of the information is stored in a database. Preliminary results indicated that implantation was affected by the number and size of blastomeres on Day 3, and prediction of embryo implantation was superior to that of traditional manual scoring systems (Paternot et al., 2009). Additional information about morphology and embryo development has also accumulated from the use of modern systems that allow time-lapse photography in combination with culture systems.



Aneuploidy Screening


Cytogenetic analysis of a biopsied polar body or blastomere has been used to screen embryos in order to detect those with an abnormal chromosome composition, a strategy known as aneuploidy screening or preimplantation genetic screening (PGS). The techniques employed for biopsy and diagnosis are described in Chapters 13 and 14, as well as the associated pros and cons. PGS continues to be a subject of considerable debate (Kuliev and Verlinsky, 2008; Mastenbroek et al., 2008; Sermondade and Mandelbaum, 2009; Dale et al., 2016; Gleicher and Orvieto, 2017): time-lapse studies have demonstrated that removing a single blastomere has a negative effect on embryo development (Kirkegaard et al., 2012b). Guidelines issued by the European Society for Human Reproduction and Embryology (ESHRE) and the American Society for Reproductive Medicine (ASRM) currently discourage the use of PGS.



Follicular Indicators of Embryo Health


Each assisted conception cycle generates a number of waste products: luteinized granulosa cells, follicular fluids harvested from follicles at the time of oocyte retrieval and cumulus cells that can be removed and separated from the oocytes. These products have all been assayed to provide indices of embryo developmental competence. Biomarkers quantified in serum and follicular fluid include cytokines, C-reactive protein and leptin (Wunder et al., 2005), inhibin B (Chang et al., 2002) and reactive oxygen species (Das et al., 2006). Cumulus cell gene expression profiles have also been linked to the implantation potential of oocytes and embryos (McKenzie et al., 2004). While many of these parameters are indicative of follicular differentiated status at the time of oocyte harvest, follicular fluid is a highly concentrated cellular exudate, which is accumulated over an extended period. Consequently, to date neither follicular fluid nor granulosa cell assays at the time of oocyte collection have provided a consistent measure for assessing the implantation potential of individual embryos. The degree of follicular vascularization and its relationship to mitochondrial segregation in embryonic blastomeres has also been promoted as a determinant of embryo developmental competence (Van Blerkom et al., 2000).



Secreted Factors


Assessment of products secreted by the embryo, the embryonic ‘secretome,’ may be a better indicator of embryo development in vitro and in vivo than measurements of the follicular environment. For example, secretion of factors that regulate gamete transport and/or prepare the female tract for implantation has been used to predict embryo health. In this context, measurement of the amount of soluble human leukocyte antigen-G (HLA-G) into embryo culture media has been directly related to embryo quality and viability (Sher et al., 2004). mRNAs for HLA-G can be detected in human blastocysts, but the cellular origins and biology of soluble HLA-G are not clear (Sargent, 2005). The suitability of HLA-G as a predictor of embryo developmental potential has also been questioned, as the amount of HLA-G measured in embryo culture media appears to exceed the total protein content of the embryo itself (Ménézo et al., 2006).



Molecular Approaches to Embryo Assessment


Analysis of medium that has been used for embryo culture offers further approaches to determine embryo ‘health’/viability. Three different types of substance have been studied as potential biomarkers:




  1. 1. Proteins translated from specific gene expression products (proteomics)



  2. 2. Products of embryo metabolism, the metabolome: amino acids, oxidation products, carbohydrates, carboxylic acids (metabolomics)



  3. 3. RNA-based regulators of gene expression (small noncoding interference RNA, micro messenger RNA [miRNA]).


Since such molecules may be found in only minute quantities, assay methods must be highly sensitive, with cost efficiency and methodology that can be applied in clinical IVF. Apart from small sample volumes and low analyte concentrations, the presence of albumin in the media also makes it difficult to detect other proteins in a 60- to 70-kDa range.


Advances in technology facilitated noninvasive measurement of amino acid uptake/output into the spent culture medium of individual embryos, and products of embryo metabolism have been quantified and used in attempts to identify healthy embryos. During early development in vivo and in vitro, each preimplantation embryo will utilize numerous substrates from its immediate environment: oxygen for respiration, sugars, energy sources such as glucose and proteins/amino acids. The embryo also secretes many waste products of metabolism into its immediate environment. The turnover of these substrates and products has been measured in embryos themselves and in their culture environment (Gardner, 2007). Development of technologies to measure embryo metabolism led to the science of ‘embryo metabolomics,’ defined as ‘the systematic analysis of the inventory of metabolites – as small molecule biomarkers – that represent the functional phenotype at the cellular level’ (Posillico et al., 2007). Amino acid turnover in spent embryo culture media has been measured noninvasively using high performance liquid chromatography (HPLC), gas chromatography and mass spectrometry, nuclear magnetic resonance spectroscopy and Raman near infrared spectroscopy. Embryo oxygen consumption by respirometry has also been used to quantify embryo metabolism. Details of the advantages and disadvantages associated with each of these different methods are reviewed by Rødgaard et al. (2015). Potential biomarkers were identified in different trials, but none have proved to be consistently helpful for clinical use.


The use of amino acid turnover to predict the developmental competence of individual embryos is based on the premise that metabolism is intrinsic to early embryo health and that the embryonic metabolome is immediately perturbed when embryos are stressed (Houghton and Leese, 2004; Lane and Gardner, 2005, Leese, 2012). Amino acid profiling has been extensively tested as a valid clinical diagnostic test for embryo selection in animal species, including mice, cows and pigs, as well as humans (Houghton et al., 2002). Collectively the data on amino acid metabolism across these species indicate that:




  1. 1. The net rates of depletion or appearance of amino acids by individual embryos vary between amino acids and the stage of preimplantation development.



  2. 2. There is no difference between the turnover of essential and nonessential amino acids as defined by Eagle in 1959.



  3. 3. The turnover of amino acids is moderated by the concentrations of the amino acids in the embryo culture media.


The association of metabolic profiles of certain amino acids (particularly asparagine, glycine and leucine) with embryo viability is based on a variety of complex interactions that involve energy production, mitochondrial function, regulation of pH and osmolarity. Nevertheless, on the basis of the turnover of three to five key amino acids, morphologically similar cleavage stage human embryos which are metabolically ‘quiet’ were identified (Leese, 2002). These were identified to have the capacity to undergo zygotic genome activation and blastocyst development, and were compared with embryos which are metabolically active but are destined to undergo cleavage arrest (Houghton et al., 2002). Interestingly, amino acid turnover by early cleavage embryos appears to be linked to embryo genetic health (Picton et al., 2010). In this context inadequate energy production has been postulated as a cause of aneuploidy induction, due to errors during the energy-dependent processes of chromosome alignment, segregation and polar body formation (Bielanska et al., 2002a; Ziebe et al., 2003).


Preliminary trials of in-vitro amino acid profiling suggested that this strategy might be used to identify cleavage stage embryos with high implantation potential (Brison et al., 2004). Metabolic profiles of developmentally competent, frozen-thawed human embryos were also consistent with those of fresh embryos (Brison et al., 2004), and metabolic profiles could be used to identify frozen-thawed embryos with the potential to develop to the blastocyst stage in vitro.


Although underlying principles and concepts may be valid and initial results were encouraging, the technical demands of the assays and complexity of diagnosis have so far restricted their clinical application.



Micro Messenger RNA (miRNA)


miRNAs are small, stable noncoding RNA molecules of 18–22 nucleotides that are thought to be negative regulators of gene expression, driving numerous cellular processes by repressing translation of mRNA. They are secreted from cells into the extracellular environment, where they may act as signaling molecules, and have been detected in urine, milk, saliva, semen and blood plasma. Their role in embryogenesis and development both in humans and in other species has been reviewed by Galliano & Pellicer (2014). The types and amounts of miRNA secreted into culture medium has been analyzed by quantitative real-time polymerase chain reaction (qPCR), with initial results suggesting that differences between viable and nonviable embryos can be detected (Kropp et al., 2014; Rosenbluth et al., 2014). The qPCR method is fast, relatively inexpensive and can be used with small sample volumes, but several logistic challenges still remain before it can be applied in routine clinical IVF.



Mitochondrial DNA (mtDNA)


Studies examining the levels of mtDNA in blastocyst trophectoderm have suggested that embryos with high levels of mtDNA do not implant, and quantification of mtDNA in blastocyst trophectoderm has been proposed as another biomarker of embryo viability (Fragouli et al., 2017; see Cecchino & Garcia-Velasco, 2019 for review). However, conflicting results continue to challenge the potential significance of mtDNA in human embryo implantation, and a number of studies have not been able to confirm this observation (Humaidan et al., 2018). Methods used to quantify mtDNA have technical limitations, in particular due to the challenge of quantification in a small number of cells, as well as the fact that the number of cells in a trophectoderm biopsy varies from sample to sample. To compensate for this factor, the amount of mtDNA detected must be normalized against reference DNA sequences in the nuclear genome, and this crucial parameter can differ between protocols/studies. Our current understanding is inadequate in many areas, and further validation of protocols is essential. The technical challenges of mtDNA quantification must be fully appreciated in order to understand the significance of study results (Wells, 2017).


High levels of mtDNA are thought to be a reflection of cellular stress, and the levels could be influenced by factors specific to their in-vitro culture, such as culture medium, handling procedures or ovarian stimulation protocols, as well as by patient-specific parameters. It is possible that the phenomenon of elevated mtDNA as an indicator of embryo stress might be usefully applied in attempts to optimize aspects of IVF treatment.



Gene Expression Studies


The activity of individual genes in an embryo change continuously in response to fluctuating intrinsic requirements or environmental conditions. Embryonic gene expression has been assessed at different stages of preimplantation development, using real-time PCR analysis of cDNA fragments as a measure of mRNA transcripts as well as by single-cell RNA-sequence profiling (Adjaye et al., 1998, Yan et al., 2013). More recently, in-vitro transcription techniques have been developed that allow oocyte and embryo mRNA to be amplified to a level that can be analyzed by microarrays (Wells, 2007). Factors related to patient or treatment have been shown to have an effect on gene expression in oocytes, morulae and blastocysts (Mantikou et al., 2016), and different patterns of gene expression have been associated with the use of different types of culture media (Kleijkers et al., 2015). Although the technology is far from routine clinical application, as a research tool it is hoped that data accumulated from gene expression studies over time may eventually lead to the identification of markers for embryo viability and implantation potential.


Overall, the search for biomarkers as predictors of developmental potential has highlighted important fundamental principles, in particular the fact that there may be differences between different patient populations, or even between individual embryos in a single cohort. It is now clear that early stage embryos have a high degree of plasticity, reflected by the effect of culture conditions on patterns of gene expression. Sophisticated analytical methods are capable of producing hundreds or thousands of variables that can mask a drift in accuracy; there are no simple ‘standards,’ and different analytical methods can produce varying results, making them difficult to standardize. Although the technologies continue to contribute to our basic knowledge about early embryo development, such concerns make the techniques difficult/impossible to introduce into clinical routine, and morphological assessment continues to be the most appropriate tool in selecting embryos for transfer in a routine IVF laboratory.



Embryo Grading


Divergent national strategies define the maximum number of embryos that can be transferred in any one cycle, and elective single embryo transfer (eSET) is recommended for selected patient populations, in order to decrease the incidence of multiple births associated with ART. In the United Kingdom, the HFEA introduced a ‘Multiple Births, Single Embryo Transfer Policy’ in 2009, setting a target multiple birth rate for each UK clinic of 10%.


In routine practice, embryo selection for transfer continues to be based primarily on morphological assessment, often with multiple observations over the course of the embryo’s development. Disturbance to the embryo culture system should be minimized, with assessment no more than once per day; some laboratories skip assessment on Day 2 or Day 4. Embryo development should also be checked within a time period that is appropriate to the laboratory workload.


Routine morphological assessment is highly subjective, with both inter- and intraobserver variation repeatedly documented. In addition, embryo development is affected by differences in culture media and environment, as well as by handling procedures. These two factors make it difficult to compare embryo scores/success rates between different units, and embryo assessment procedures should be validated both within laboratories and for each individual carrying out the assessment. Online external quality control schemes provide an extremely useful tool for training and validation procedures.


For cleavage stage embryos, assessment criteria include:




  • rate of division judged by the number of blastomeres



  • size, shape, symmetry and cytoplasmic appearance of the blastomeres



  • presence of anucleate cytoplasmic fragments



  • appearance of the zona pellucida.


Criteria are frequently combined to produce composite scoring systems which may incorporate pronuclear scoring of zygotes. Online quality control schemes are available, where participants can compare their scoring criteria for embryo assessments with those of others (www.fertaid.com, www.embryologists.org.uk/).


Although morphological assessment is recognized to be highly subjective, arbitrary and unsatisfactory, it is quick, noninvasive, easy to carry out in routine practice, and does help to eliminate those embryos with the poorest prognosis. Evaluation of blastomere shape, size and number will reflect synchronous cleavage of the blastomeres, and embryos with asynchrony in either the timing of cleavage or the process of blastomere division will be given lower scores. Unfortunately embryo cleavage in vitro rarely follows the postulated theoretical timing of early development, and computer-assisted morphometric analyses confirm that large variations in blastomere size and fragmentation are frequently observed; large variations in blastomere size have been linked to increased chromosomal errors (Hnida and Ziebe, 2007).



Zygote Scoring


Schemes for identifying healthy viable embryos at the zygote stage were proposed by Scott et al. (2000) and Tesarik and Greco (1999). Criteria suggested as predictive of optimal implantation potential include:




  • close alignment of nucleoli in a row



  • adequate separation of pronuclei



  • heterogeneous cytoplasm with a clear ‘halo’



  • cleavage within 24–26 hours.


The scoring systems were reviewed by James (2007) and are compared in Table 11.1. The timing of assessment is critical, as pronuclear development is a dynamic process, and zygote scoring should therefore be used with caution and only in conjunction with other methods of evaluation.




Table 11.1 Comparison of pronuclear morphology scoring systems, with a representative illustration of pronuclei in each scoring group

















































































Scott et al., 2000 Series 1 Scott et al., 2000 Series 2 Scott, 2003 Tesarik and Greco, 1999 Tesarik et al., 2000 Pronuclear morphology
Grade 1 Z1 Z2 Pattern O Pattern O
Grade 3 Z2 Z2 Pattern O Pattern O
Grade 2 Z3 Z3–2 Pattern 2 Non-pattern 0
Grade 4 Z3 Z3–1 Pattern 5 Non-pattern 0
Grade 4 Z3 Z3–4 Pattern 1 Non-pattern 0
Grade 4 Z3 Z3–4 Pattern 3 Non-pattern 0
Grade 5 Z3 Z3–1 Pattern 4 Non-pattern 0
Grade 5 Z4 Z4–2 Pattern 4 Non-pattern 0
Grade 5 Z4 Z4–1 NA Non-pattern 0


Early Cleavage


Timing of the first cell division of the embryo has been investigated as a predictor of developmental competence, with the suggestion that ‘early cleavage’ is associated with higher pregnancy rates. However, there is only a certain window where true early cleavage can be seen, and this extra assessment may be difficult to fit into the normal routine of an IVF laboratory. Once the embryos have undergone the second division to the four-cell stage, those that might have cleaved a few hours earlier will have similar morphology to those that did not, and it is not possible to differentiate between them.



Multinucleation


Multinucleated blastomeres can sometimes be observed at early cleavage stages, most easily on Day 2 (Figure 11.5). Studies have revealed that blastocysts developing from embryos showing multinucleation at early cleavage stages have similar aneuploidy rates to blastocysts that resulted from embryos that did not have multinucleation. Transfer of binucleated and multinucleated frozen-thawed embryos does not apparently increase the incidence of congenital anomalies and chromosomal defects in newborns (Seikkula et al., 2018). Aneuploidy activates a spindle-apparatus checkpoint in different types of cells (Wenzel & Singh, 2018), and it is possible that blastomere multinucleation may represent the activation of a cell cycle checkpoint that can convert a mosaic embryo to one that is euploid (Tesarik, 2018).





Figure 11.5 A two-cell embryo showing two nuclei in each blastomere.


Reprinted with permission from Meriano et al. (2004).


Cumulative Scoring Systems


Multi-day scoring systems can enhance embryo selection by combining both developmental rate and morphological assessment (Skiadas and Racowsky, 2007). These should provide a more accurate picture of developmental progression than can be obtained from a single observation. However, the ultimate combination of morphological features required for optimal evaluation of developmental competence has yet to be resolved. The optimal timing of embryo transfer will be more accurately determined when agreement on this is reached. In addition, multi-day assessment requires repeated embryo handling and exposure to potentially hazardous conditions outside of the incubator (see section ‘Time-Lapse Systems (TLS)’).



Fragments


Fragmentation in the human embryo is very common, affecting up to 75% of all embryos developed in vitro (Alikani, 2007); it is not clear whether this is an effect of culture conditions and follicular stimulation, or a characteristic of human development (Figure 11.6). Extensive fragmentation is known to be associated with implantation failure, but the relationship between the degree of fragmentation and the developmental potential of the embryo is far from clear. Alikani et al. (1999) found that when embryos with more than 15% fragmentation were cultured to blastocyst stage, they formed fewer morulae, fewer cavities and fewer blastocysts compared to those embryos with less than 15% fragmentation. When fragmentation was greater than 35%, all processes were compromised. Retrospective analysis of embryo transfer data revealed that nearly 90% of embryos selected for transfer were developed from embryos with less than 15% fragmentation observed on Day 3.


Figure 11.6



(a) Day 3 human embryo with type 3 fragments.





(b) Day 3 embryo with type 4 fragmentation.


Alikani and Cohen (1995) used an analysis of patterns of cell fragmentation in the human embryo as a means of determining the relationship between cell fragmentation and implantation potential, with the conclusion that not only the degree, but also the pattern of embryo fragmentation determine its implantation potential. Five distinct patterns of fragmentation which can be seen by Day 3 were identified:




  • Type I: <5% of the volume of the perivitelline space (PVS) occupied by fragments.



  • Type II: small, localized fragments associated with one or two cells.



  • Type III: small, scattered fragments associated with multiple cells.



  • Type IV: large, scattered fragments associated with several unevenly sized cells and scattered throughout the PVS.



  • Type V: fragments throughout the PVS, appearing degenerate such that cell boundaries are invisible, associated with contracted and granular cytoplasm.


Some embryos have no distinct pattern of fragmentation.


No definite cause of fragmentation has been identified, although speculations include high spermatozoal numbers and consequently high levels of free radicals, temperature or pH shock, and stimulation protocols. Apoptosis has been suggested as a possible cause, and a progressive shortening of telomere length, which induces apoptosis, has been linked with fragmentation (Keefe et al., 2005); however, this study was not definitive. Mitotically inactive cells do not exhibit fragmentation (Liu et al., 2002), and it has been suggested that aberrant cytokinesis in the presence of spindle and cytoskeletal abnormalities may be associated with fragmentation.


Observed via the scanning electron microscope, the surface of fragments is made up of irregularly shaped blebs and protrusions, very different to the regular surface of blastomeres, which is organized into short, regular microvilli (Figure 11.7).





Figure 11.7 Scanning electron micrographs. (a, b) Two views of a human four-cell embryo showing 20% fragmentation. (c–e) Magnification of corresponding areas showing regular short microvilli of vital blastomeres and intercellular areas. (f) Magnification of the surface of a cytoplasmic fragment showing irregular blebs and protrusions.


Reproduced with permission from Dale et al. (1995).

Interestingly, programmed cell death in somatic cells also starts with surface blebbing, and is caused, in part, by a calcium-induced disorganization of the cytoskeleton. We can speculate that similar mechanisms operate within human embryos, but there is so far no scientific evidence that this is the case.


There does appear to be an element of programming in this partial embryonic autodestruction, as embryos from certain patients, irrespective of the types of procedure applied in successive IVF attempts, are always prone to fragmentation. Surprisingly, fragmented embryos, repaired or not, do implant and often come to term. Time-lapse photography technology has clearly demonstrated that an individual embryo can radically change its morphological appearance in a short period of time: fragments that are apparent at a particular moment in time can be subsequently absorbed with no evidence of their prior existence (Hamberger et al., 1998; Hashimoto et al., 2009; Mio and Maeda, 2008). This demonstrates the highly regulative nature of the human embryo, as it can apparently lose over half of its cellular mass and still recover, and also confirms the general consensus that the mature oocyte contains much more material than it needs for development. The reasons why part, and only part, of an early embryo should become disorganized and degenerate are a mystery. Different degrees of fragmentation argue against the idea that the embryo is purposely casting off excess cytoplasm, somewhat analogous to the situation in annelids and marsupials that shed cytoplasmic lobes rich in yolk, and favor the idea of partial degeneration. Perhaps it involves cell polarization, where organelles gather to one side of the cell. It is certain that pH, calcium and transcellular currents trigger cell polarization, which may in certain cases lead to an abnormal polarization, and therefore to fragmentation. Describing fragmentation as a degenerative process may not be justified, but more research is needed to elucidate whether the implantation of embryos with extensive fragmentation at the cleavage stages has any long-term effects.



Embryo Grading: Cleavage Stages


Despite the introduction of extended culture and sophisticated monitoring and genetic screening technologies, a significant number of apparently high-quality and chromosomally normal embryos still do not implant, and selecting embryos for transfer remains an ongoing challenge in routine clinical IVF. After more than 40 years, and the birth of close to 10 million babies, the basic selection criteria in a routine IVF laboratory without research facilities continue to be based on assessment of morphology, despite numerous attempts to find new biomarkers and variables (Figure 11.8). Preimplantation genetic screening has shown a surprising discrepancy between gross morphology and genetic normality of embryos. Even the most ‘beautiful’ Grade 1 embryos may have numerical chromosomal anomalies, whilst those judged to be of ‘poorer’ quality, with uneven blastomeres and fragments, may have a normal chromosome complement. The variables that continue to prove consistently useful as predictors of viability remain: numbers of cells and fragmentation, cell symmetry and nucleation. (See Van den Bergh et al. [2012] for an atlas of >1000 embryo images together with their clinical/IVF histories and cycle outcomes.)






(a) Two-cell embryo on Day 2, no fragments.





(b) Day 2 embryo, Grade 3.





(c) Day 2 embryo, uneven blastomeres with one large dominant blastomere.





(d) Day 2 embryo, one blastomere shows large vacuole, Grade 2/3.





(e) Four-cell embryo on Day 2, Grade 1.





(f) Grade 3 embryo on Day 2.





(g) Grade 4 embryo on Day 2.





(h) Day 3 six-cell embryo, Grade 1.





(i) Day 3 eight-cell embryo, Grade 1.


Images (c) and (g) kindly supplied by Thomas Ebner, Austria; (g) by Marc van den Bergh, Switzerland; (e) and (f) by Oleksii Barash, Ukraine; (h) by Gemma Fabozzi, Naples. Despite the extreme difference in their morphology, the single transfer of embryos (e) and (f) both led to healthy live births (photos and history courtesy of Oleksii Barash, Ukraine).

Only gold members can continue reading. Log In or Register to continue

Stay updated, free articles. Join our Telegram channel

Sep 17, 2020 | Posted by in OBSTETRICS | Comments Off on Chapter 11 – Oocyte Retrieval and Embryo Culture

Full access? Get Clinical Tree

Get Clinical Tree app for offline access